Next Article in Journal
The Roles of Various Prostaglandins in Fibrosis: A Review
Next Article in Special Issue
Sustainable Agriculture Systems in Vegetable Production Using Chitin and Chitosan as Plant Biostimulants
Previous Article in Journal
Attacking COVID-19 Progression Using Multi-Drug Therapy for Synergetic Target Engagement
Previous Article in Special Issue
Transcriptional Cascade in the Regulation of Flowering in the Bamboo Orchid Arundina graminifolia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Improving Performance of Salt-Grown Crops by Exogenous Application of Plant Growth Regulators

1
Tasmanian Institute of Agriculture, University of Tasmania, Prospect 7250, Australia
2
International Research Centre for Environmental Membrane Biology, Foshan University, Foshan 528000, China
3
College of Agronomy, Shanxi Agricultural University, Taigu 030801, China
*
Author to whom correspondence should be addressed.
Biomolecules 2021, 11(6), 788; https://doi.org/10.3390/biom11060788
Submission received: 12 March 2021 / Revised: 19 May 2021 / Accepted: 21 May 2021 / Published: 24 May 2021

Abstract

:
Soil salinity is one of the major abiotic stresses restricting plant growth and development. Application of plant growth regulators (PGRs) is a possible practical means for minimizing salinity-induced yield losses, and can be used in addition to or as an alternative to crop breeding for enhancing salinity tolerance. The PGRs auxin, cytokinin, nitric oxide, brassinosteroid, gibberellin, salicylic acid, abscisic acid, jasmonate, and ethylene have been advocated for practical use to improve crop performance and yield under saline conditions. This review summarizes the current knowledge of the effectiveness of various PGRs in ameliorating the detrimental effects of salinity on plant growth and development, and elucidates the physiological and genetic mechanisms underlying this process by linking PGRs with their downstream targets and signal transduction pathways. It is shown that, while each of these PGRs possesses an ability to alter plant ionic and redox homeostasis, the complexity of interactions between various PGRs and their involvement in numerous signaling pathways makes it difficult to establish an unequivocal causal link between PGRs and their downstream effectors mediating plants’ adaptation to salinity. The beneficial effects of PGRs are also strongly dependent on genotype, the timing of application, and the concentration used. The action spectrum of PGRs is also strongly dependent on salinity levels. Taken together, this results in a rather narrow “window” in which the beneficial effects of PGR are observed, hence limiting their practical application (especially under field conditions). It is concluded that, in the light of the above complexity, and also in the context of the cost–benefit analysis, crop breeding for salinity tolerance remains a more reliable avenue for minimizing the impact of salinity on plant growth and yield. Further progress in the field requires more studies on the underlying cell-based mechanisms of interaction between PGRs and membrane transporters mediating plant ion homeostasis.

1. Introduction

Salinity is one of the major abiotic stresses affecting crop plants and limiting production worldwide [1,2]. Globally, approximately 1125 million ha of land is affected by soil salinity [3]. Soil salinization is increasing at a rate of ~3 ha/min [4], and now becoming a major concern for the irrigated agriculture [1,5]. Salinity stress induces a multitude of responses in plants at various levels of plant structural organization. The three primary constraints imposed by salinity on plants are osmotic stress, ionic disbalance/toxicity, and oxidative stress [1]. Osmotic stress decreases external water potential and leads to a reduced water uptake capacity of plants, thus affecting cell expansion growth. It also leads to stomata closure, reducing the plant’s ability to assimilate CO2. The ionic stress is caused by an excess uptake of toxic salt ions (mainly Na+ and Cl) that hamper normal metabolic processes in plants. The accumulation of toxic Na+ and Cl is also accompanied by a massive reduction in cytosolic K+, with numerous implications for a cell’s metabolic activity and viability [6,7,8,9]. Cl toxicity is less drastic compared with Na+, but nonetheless can cause a significant disturbance to many physiological and biochemical processes in sensitive species [10,11]. Salinity stress also leads to the production of excess amounts of reactive oxygen species (ROS) in plant tissues [12,13,14], including superoxide anion (O2−), hydrogen peroxide (H2O2), singlet oxygen (1O2), and hydroxyl radical (OH). These ROS can severely damage the plant’s cellular structures and macro molecules like DNA, enzymes, and lipids [15,16,17,18]. Also, ROS are highly potent regulators of a broad range of Ca2+, Na+, and K+—permeable ion channels [19,20,21]—thus causing a major disturbance to stress signaling and intracellular ion homeostasis, well before damaging effects become evident.
Three major cereal crops, namely wheat, rice, and maize, are responsible for over 50% of daily caloric uptake by the human population. All of them are classified as salt-sensitive species and perform poorly when grown on saline soils. For example, wheat provides about 20% of human food energy requirements and 25% of proteins consumed daily worldwide (Wheat Initiative, www.wheatinitiative.org/, accessed on 23 May 2021). Wheat is a salt-sensitive glycophyte [22], and salinity is considered to be a major soil constraint in the Australian Wheatbelt [23,24] which results in about a 40% yield reduction [25], costing Australian economy ~A$200 million per annum [26].
Two different (but potentially complementary) approaches can be used to reduce the negative impact of salinity stress on plant growth and yield. The first one is the development of salt-resistant cultivars via molecular or classical breeding. The second approach is related to agronomical means, and includes inoculating seeds with halotolerant plant growth promoting rhizobacteria (PGPR) or the application of various plant growth regulators (PGRs) [27,28,29,30,31,32,33,34]. While genetic improvement is considered as the best solution from a long-term perspective, no significant progress has been made in breeding programs. This is due to the polygenic nature of tolerance, which reflects the complexity of salt tolerance mechanisms in plant, and the lack of available genes that confer salt stress tolerance [35,36,37]. Conventional breeding techniques are time-consuming and laborious, and have met with only a limited success. With the advancement of science and technology, molecular techniques and transgenic technology have been widely used in plant breeding worldwide. Although transgenic technology is considered as a fast and effective method to obtain salt-tolerant varieties, the public acceptance of genetically modified (GM) crops remains a major stumbling block in most countries [38,39,40]. In this context, PGPRs could potentially minimize the detrimental effects of salinity stress on plant growth and yield without triggering these public/governmental concerns. PGRs share the common function of regulating intrinsic hormone levels within plants by modulating signaling within various hormone transduction pathways, and are widely available and easy to apply to crops [41,42]. However, the effectiveness of PGPRs depends upon their interaction with host plant and soil environment. Sometimes plant growth-promoting bacteria has exhibited harmful effects on the growth and development of plants [43], and is often considered to be “unsafe” for human and animal health [44]. Also, PGRs cross-talk with each other and may act synergistically or antagonistically to regulate plant growth, development, and defense responses, generally by inducing gene expression [45]. This complexity may result in a certain level of unpredictability and negate expected beneficial effects. The main aim of this review was to summarize the bulk of the reported data on the use of PGRs for improving performance of plants grown on salt-affected lands, revealing underlying cellular mechanisms and downstream targets, and critically assessing the applicability of PGR for sustainable crop production under conditions of soil salinity.

2. Plant Growth Regulators

Plant growth regulators (PGRs) are defined as synthetic or naturally occurring organic compounds that influence biological process in higher plants at very low concentrations [41]. PGRs can increase or decrease growth and development by altering their normal biological processes [41,46,47]. When these compounds occur naturally inside the plant they are known as phytohormones, but when applied exogenously they are called PGRs [48]. PGRs act as signaling agents, allowing plants to maintain plasticity during growth and development, and are hence considered as principal factors in responses of plants to biotic and abiotic stresses [49,50]. PGRs play a significant role in alleviating salt stress via a broad range of physiological and developmental alterations [51,52,53]. In broad terms, PGRs are involved in increasing the physiological availability of water and essential nutrients, while helping plants reduce toxic salt load [54]. They also have a major impact on the antioxidant enzyme activities in plants [55,56,57,58,59,60,61,62]. PGRs induce salt tolerance by increasing the activity of ROS scavenging enzymes to maintain the ROS at a nontoxic level under stress conditions [63,64]. The ameliorating ability of PGRs depends on environmental factors that affect their absorption, the concentration at which they are applied, and the physiological state of the plant [65,66,67]. PGRs are classified into a number of distinct classes, such as auxins (Aux), cytokinins (CKs), nitric oxide (NO), brassinosteroids (BRs), gibberellins (GAs), salicylic acid (SA), abscisic acid (ABA), jasmonates (JAs), and ethylene. The following sections examine the efficiency and modes of ameliorating the effects those PGRs have on how plants respond to salinity stress.

3. Effects of Plant Growth Regulators on Plant Performance under Saline Conditions

3.1. Auxins

Auxin is a widely used plant growth regulator of low molecular weight with an aromatic ring structure [68]. Auxin is produced in growing shoot tips and transported down the main stem via the polar auxin transport (PAT) mechanism [69]. It is involved in cell division and elongation, organogenesis, and apical dominance [70,71,72,73,74]. Auxin stimulates cell elongation via increasing wall extensibility and participates in the regulation of cell wall properties by inducing wall loosening [75]. Auxins also stimulate H+–ATPase operation, creating a driving force for inorganic ion uptake that contributes to increased cell turgor [76,77].
Salt stress affects the growth of both primary and lateral roots [78,79,80]. The developmental plasticity of the plant root under saline conditions is regulated by auxin [81], and the inhibition of root growth is associated with reduced auxin accumulation [82], most likely via the PAT mechanism [83]. Slowing root growth might be an adaptive mechanism for plants surviving in salt environments. Exogenous application of auxin leads to an increase in root growth under saline conditions in several species [82,84]. However, this increase in root growth comes with a carbon cost, and therefore, could be counterproductive under conditions of severe salinities, when (limited) plant ATP pull is required for a range of defense responses. Lateral root development is also an important survival strategy for plant to avoid damage in unfavorable environmental conditions, including salt stress [85,86,87]. Plants remodel their root architecture by altering auxin accumulation and its redistribution under salt stress conditions [88,89,90,91].
Indole-3-acetic acid (IAA) is an important member of auxin family of plant hormones [92,93]. Stress conditions lead to a significant reduction in IAA concentrations in rice, maize, tomato, and wheat plants [94,95,96,97]. This reduction of free IAA concentration in crop is cultivar- or plant organ-dependent. For example, the free IAA concentration remained constant in the roots of a salt-tolerant maize hybrid line, but significantly decreased in the roots of a salt-sensitive line, while IAA concentration in leaves remained constant in both lines [95]. Therefore, the reduction in plant growth and development under stress conditions could be an outcome of altered auxin accumulation and redistribution. Consistent with this, some studies suggested that exogenous application of Aux may alleviate salinity stress in many crops, including wheat [45,84,98,99]. For example, foliar spraying of 2 mM IAA to salt-grown maize plants increased kernel yield by ~9%, while a combination of IAA spray and a basal application of inorganic nutrients (K and P) improved yield by 20% under 10 dS m−1 saline conditions [98]. In rice, a foliar spray of IAA (50 mL per pot) at the reproductive stage increased grain yield by ~4% and ~46% in salt-tolerant and salt-sensitive cultivars, respectively, under 6 dS m−1 saline conditions [99]. In the case of wheat, seed primed with 100, 150, and 200 mg L−1 of different synthetic auxins, namely 4-dichlorophenoxyacetic acid (2,4-D); 2,4,5-trichlorophe-noxyacetic acid (2,4,5-T); and α-naphthaleneacetic acid (NAA) improved the performance of salt-grown plants under both controlled environment and field conditions. However, the effect was cultivar-, priming agent-, and concentration-dependent, questioning the direct causal link between PGR application and observed responses. Seeds primed with 150 mg L−1 NAA increased plant grain yield by ~36% and ~27% in salt-sensitive and salt-tolerant wheat varieties, respectively, when grown under 150 mM saline conditions [100]. The use of IAA-producing bacteria also lead to an improvement of wheat performance under saline conditions [101].
One of the modes of action for the auxin-driven decrease in the extent of salt damage on plant performance is increasing the activities of auxin-responsive genes that increase the plant’s water retention capacity and decrease H2O2 accumulation [45,102]. The auxin responsive TaIAA gene family exhibits differential expression during the absorption of inorganic salts in wheat roots [103], and an auxin responsive gene small auxin-upregulated RNAs TaSAUR75 in wheat regulates plant growth and development under saline conditions by preventing H2O2 accumulation under salt stress [102]. The salt-responsive gene TaEXPB23 is downregulated by the exogenous application of IAA, and overexpression of TaEXPB23 conferred salt stress tolerance by decreasing osmotic potential and enhancing the water retention ability in transgenic tobacco [45].

3.2. Cytokinins

Cytokinins (CKs) are an important class of phytohormones that promote cell division in the roots and shoots of plants. There are two types of CKs: the adenine type and the phenylurea type [104]. To date, phenylurea CKs have not been found in plants [105]. The adenine-type plant hormones are the main form and classified as isopentenyladenine (iP), trans-zeatin (tZ) and cis-zeatin (cZ), and dihydrozeatin and its riboside [106,107]. The iP-type cytokinin in the phloem is transferred from aerial parts of the plant to the root to maintain vascular structure in the root meristem [108]. The tZ-type cytokinin is transferred from the root to aerial parts of the plant through the xylem to regulate shoot growth [109]. As a result of these redistributions, CKs participate in various biochemical and physiological processes, such as cell division and leaf senescence, thus controlling the root/shoot ratio. The long-distant CK transport is also essential for plant responses to abiotic stresses, including salt [110,111] (Figure 1).
CK has both a positive and negative regulatory role in alleviating the detrimental effects of salinity stress. For example, cytokinin deficiency leads to enhanced salinity tolerance in Physcomitrella patens and Arabidopsis [111,112,113], and contributed to yield improvement in many crops [114]. Cytokinin oxidase (CKX) is the prime enzyme involved in CK metabolism, which can effectively reduce the concentration of CK in plants. The overexpression of CKX significantly influences hyposensitivity to salt stress in Physcomitrella patens [111,112]. Regulation of isopentenyltransferase (IPT) genes under conditions of salinity stress also reduces CK content in Arabidopsis and enhances its tolerance to salinity [112,115,116]. On the other hand, overproduction of CKs by inducible expression of IPT8 gene leads to reduced adaptive ability in Arabidopsis. The excess production of CKs decreases transcript levels of ROS scavenging enzymes, thus leading to increased ROS production that correlates with sensitivity to salt stress [113]. In contrast to the above negative effects of CKs, other studies have indicated a beneficial role of CK for plant performance under salinity stress. The application of CKX inhibitor INCYDE protected the photosynthetic apparatus and increased the production of flowers in tomato plants [117]. The downregulation of CKX2 under saline stress condition significantly increased CK concentration and reduced yield penalty in rice [118,119]. AGO2 (argonaute RISC catalytic component 2) plays a key role in improving salinity tolerance by changing the level of CKs and enhancing grain yield in rice [120]. These contradictory reports question the practical application of CKs, and suggest concentration- and tissue-specific modes of action.
Published papers indicate that salt stress conditions are often associated with a decrease in CK concentration in crops like rice and wheat. CK spraying leads to an increase in grain yield under saline conditions [121]. Exogenous CKs may increase plant antioxidant enzymes activity and reduce ROS load. Kinetin (adenine-type CKs) spraying at a rate of 10 mg L−1 ameliorated the deleterious effects of salinity by reducing the uptake of toxic ions Na+ and Cl and promoting the uptake of K+ in wheat seedlings. Seed primed with different concentrations of synthetic cytokinins (kinetin and benzylaminopurine) increased grain yield by up to 52%, but in a strong cultivar- and concentration-dependent manner [60].
CKs mainly promote physiological responses through the regulation of gene expression [122,123], but little information is available about the molecular function of cytokinin under salt stress conditions. In Arabidopsis, CRF6 (cytokinin response factor 6) represses cytokinin-associated genes during oxidative stress [124]. A salt-inducible novel wheat TaCKX3 (cytokinin oxidase/dehydrogenase) gene is located on chromosome 7B [125], and silencing of the TaCKX1 gene increased grain yield in wheat [126]. The expression of the high-affinity potassium transporter AtHKT1.1, which controls xylem Na+ loading, was repressed by CK treatment in Arabidopsis [127]. The genes involved in ROS breakdown are also greatly affected in cytokinin-deficient mutant ipt1,3,5,7 [113]. CKs help induce cytokinin response factors (CRFs) in the ERF-VI subfamily; CRFs positively regulate osmotic stress tolerance [128,129]. However, other reports have suggested that overproduction of CK resulted in a negative effect in plants by modulating stress-responsive gene expression. For example, overexpression of cytokinin biosynthetic gene AtIPT8 (adenosine phosphate–isopentenyl transferase 8) significantly inhibits true leaf emergence and primary root growth under salt stress conditions. It is also associated with increasing ROS production, decreasing survival rates, and chlorophyll content, which lead to reduced salinity tolerance [113]. These pleiotropic effects question the practicalities of CK application and suggest that balancing CK levels is essential for adapting plants to salt conditions.

3.3. Nitric Oxide

Nitric oxide (NO) is a gaseous free radical that acts as a signaling molecule. NO synthesis in plants is mainly carried out by L-arginine-dependent, nitric oxide synthase-like activity and nitrate reductase (NR)-catalyzed reduction of nitrite (NO2) [135,136,137,138]. It is widely known as a “jack-of-all-trades” in stress responses [139,140]. Exogenous application of NO enables plant protection against various abiotic stresses, including salinity [141,142,143,144]. For example, exogenous NO treatment increases K+ concentration and decreases Na+ concentration in salt-grown plants, thus maintaining an optimal K/Na ratio that is critical for plant’s operation [145]. NO alleviates osmotic stress by scavenging reactive oxygen species through increased antioxidant enzyme activity [141,146,147,148] and glucose-mediated repression of photosynthesis [148]. NO protects the mitochondria from oxidative damage by increasing ATP synthesis, and seed priming with NO increases wheat grain yield by up to 22% under saline conditions [147]. NO also significantly modulates both H+–ATPase and H+–PPase (H+–pyrophosphatase) activities, thus conferring salinity tolerance in plants [149]. These results suggest that application of exogenous NO could potentially improve crop growth and development under salt stress conditions.
Despite the above beneficial reports, the practical application of NO in ameliorating the detrimental effects of salinity is questionable for several reasons. First, similar to the cytosolic Ca2+ signaling, stress-induced elevation in NO levels is usually transient and requires a return to the basal level. Second, NO is involved in multiple signaling pathways, so alteration in basal NO level caused by its exogenous application may interfere with some of them. Last but not least, the biological lifetime of an NO molecule is relatively short (millisecond range [150]), questioning its ability to sustain long-term control of transporters activity.

3.4. Gibberellins

Gibberellins belongs to a large group of tetracyclic diterpenoid carboxylic acid derivatives which have various physiological functions, such as stimulating organ growth through enhancement of cell division and cell elongation [151,152]. Gibberellic acid (GA) is the most common form of gibberellin [153,154,155]. The biosynthesis of GA is regulated by both developmental and environmental stimuli [156]. Salinity stress reduces endogenous GA content, resulting in plant’s hypersensitivity to salt [157,158].
DELLA family protein is a major GA-negative regulator that may be involved in different environmental and hormonal signaling. For example, DELLA protein SLR1 plays a role in inhibiting plant growth by inhibiting GA signaling under salt stress conditions [159]. Overexpression of some other GA catabolism-related genes like OsGA2ox5 [160] and OsMYB91 [161] in rice and AtGA2ox7 [162] in Arabidopsis reduces growth and shows an enhanced tolerance to salt stress compared to wild plants. Again, OsCYP71D8L is a potential GA-deactivating protein that plays a significant role in balancing the growth process and stress responses, and leads to enhanced tolerance to salt stress in rice [163]. These results suggest that the reduction of GA signaling under salt stress conditions is directly associated with salt tolerance in plants. On the other hand, some papers have reported that exogenous application of GA had a positive effect on salt stress tolerance in many crops. In this context, increasing lipid biosynthesis is one of the essential mechanisms conferring salinity stress tolerance in plants. This process is disturbed by salinity, but the exogenous application of GA leads to up-regulation of chloroplast lipid biosynthesis, which is directly associated with increasing salt stress tolerance in rice [164]. Salt stress also reduces enzyme activities, as well as hampering the nutritional balance in plants. A foliar spray of 0.1 mM GA significantly alleviates the damaging effect of salt and increases growth and enzymatic activities in okra [165]. Another potential target of GA are expansins, which determine extensibility and mechanical properties of cell walls. The TaEXPB23 transcript expression in wheat was upregulated by salt stress but downregulated by exogenous GA application, and constitutive overexpression of TaEXPB23 enhanced salt stress tolerance in transgenic tobacco by enhancing water retention ability and decreasing osmotic potential [45].

3.5. Brassinosteroids

Brassinosteroids are primarily polyhydroxylated, sterol-derived plant growth regulators. They are ubiquitous in all plant species, and are implicated in a wide range of growth and developmental processes in various crop plants [166,167,168,169,170]. BRs are involved in the regulation of multiple physiological, developmental, and biochemical processes, including seed germination, cell division and elongation, differentiation of vascular tissues, development of root and shoots, senescence, reproduction, and photomorphogenesis. BRs are also essential for plant adaptation to various abiotic and biotic stresses [166,171,172]. BRs interact with other hormones to regulate these types of activities in plants [169]. The most common, and thus best studied, are 24-Epibrassinolide (24-EBL) and 28-homobrassinolide (28-HBL) [173,174]; studies of these have included their role in mitigating salt stress in plants [175,176].
A foliar spraying of EBL ~10−8 M or seed soaking with ~10−6 M scavenged excessive ROS through the enhancement of antioxidant enzyme activities and modified the activity of proline metabolism, thus improving salinity stress tolerance in wheat [177]. Exogenous 5 μM EBL treatment resulted in a ~42% increase in pod yield in salt-grown bean plants [178], and foliar spray of 0.125 mg L−1 BR led to 18–35% increase in seed yield in peas [179].
The molecular mechanisms explaining how BRs control stress responses and regulate stress-responsive gene expression in plants are largely unknown [180,181,182,183]. BRs bind to a small family of leucine-rich repeat receptor kinases (BRI1) at the cell surface, thereby initiating an intracellular signal transduction cascade that results in altered expression of hundreds of genes that are implicated for diverse functions, including increased adaptation to various stresses [184]. For example, enhancing BR signaling activity in Arabidopsis led to increased salt stress tolerance, but BR-defective mutants showed sensitivity to salt stress [181]. The transcript levels of the brassinosteroid receptor (OsBRI1) were greatly influenced by EBL and its combination with salt stress in rice. On the other hand, the salt responsive gene (SalT) was negligibly expressed by the combination of salt and EBL [185]. The Arabidopsis ubiquitin-conjugating enzyme, UBC32, a stress-induced functional ubiquitin conjugation enzyme, is associated with endoplasmic reticulum protein degradation (ERAD) and brassinosteroid mediated growth promotion, as well as salt stress tolerance [186]. Brassinosteroids may also enhance abiotic stress tolerance through their interaction with other plant hormones, such as ABA [187,188,189]. The crosstalk between BR and ABA occurs after BR perception, but at or before BIN2, so a large portion of BR responsive genes are also regulated by ABA [190]. The specific details of this interaction need to be investigated in future studies.

3.6. Salicylic Acid

Salicylic acid is a phenolic compound and important endogenous growth regulator that participates in the regulation of biotic and abiotic stress responses in plants [191,192,193]. SA has been shown to affect membrane permeability, bud growth, growth rate, stomatal closure, mitochondrial respiration, material transfer, photosynthesis, and ion absorption. Because of this, SA plays an essential role in mediating plants’ adaptive responses to salinity [52,194], and controls membrane permeability (hence, ion uptake and transport) [195]. It also maintains redox homeostasis in cells under salt stress conditions [196,197,198] (Figure 2). External application of SA was found to improve tolerance to salt stress in wheat plants as a result of the upregulation of transcript levels of GPX1, GPX2, DHAR, GR, GST1, GST2, MDHAR, and GS, and further enhancement in the enzyme activities of AsA-GSH cycle [199]. SA also reduces the extent of oxidative damage caused by salt stress by enhancing the activities of peroxidase and catalase, as well as the production of osmoprotectant compounds, such as proline, betaine and glycine [193,200]. SA regulates the transcript levels of the genes encoding ASA–GSH cycle enzymes, such as DHAR (dehydroascorbate reductase), GPX (glutathione peroxidase), GR (glutathione reductase), GST (glutathione-S-transferase), MDHAR (monodehydroascorbate reductase), and GS (glutathione synthetase) [199]. SA treatment also improves the K+/Na+ ratio in salt-grown plants [59,201].
The effectiveness of exogenous SA at mitigating salt stress damage depends on crops and the concentration of NaCl in the growing media. For example, it can reduce the negative effect of salt stress when plants are exposed to moderate stress (0.3% and 0.6% NaCl), but cannot counteract severe salt stress (0.9%) in Caryophyllaceae [202]. Both arial spraying and seed priming with 1 mM SA have improved grain yield of wheat and pearl millet by ~13–14% [203]. The effectiveness of SA also varies with the concentration of endogenous or exogenous SA. The higher accumulation of endogenous SA led to hypersensitivity to NaCl [204,205,206], while activating the SA signaling pathway significantly improves salinity tolerance. For example, overexpression of SA receptors (MhNPR1 or AtNPR1) enhanced tolerance to salt/osmotic and oxidative stress by increasing SA signaling in tobacco [207,208]. On the other hand, a lack of SA receptor enhanced salt sensitivity in plants [209]. Similarly, exogenous application of three levels of SA (0.5, 1.0, and 1.5 mM) as a priming agent was evaluated in mung bean grown under different salt concentrations (3, 6, and 9 dS m−1). In most of the cases, a moderate concentration (1 mM SA) gave the best result in terms of ion content, gas exchange parameters, and chlorophyll content in leaves [210]. In Arabidopsis, the inhibitory effect of high salinity was exaggerated by >100 μM SA treatments, while plants benefited from <50 μM treatment during the seed germination [211].

3.7. Abscisic Acid

Abscisic acid is known as a stress hormone that mediates different types of biological and non-biological stress in plants [216,217,218]. ABA is synthesized in all de novo plant parts such as roots, flowers, leaves, and stems [219]. As an endogenous signaling molecules ABA enables plants’ survival under adverse environmental conditions, including salinity [220,221] (Figure 3).
During salt stress conditions, endogenous levels of ABA increase, which enhances plant adaptation to salinity by limiting ROS accumulation [222]. Higher accumulation of ABA also induces stomata closure, thus reducing transpiration for better water saving under osmotic stress conditions caused by salinity [223,224]. The accumulation of ABA occurs more rapidly in roots than leaves [95,225], and biosynthesis of ABA is associated with lateral root development in plants under salt stress conditions. This process is believed to be related to ABA regulation of auxin distribution under NaCl treatment. Interestingly, ABA biosynthesis inhibitor fluridone and an ABA biosynthesis mutant (vp14) successfully rescued the Arabidopsis phenotype under saline conditions [226]. Exogenous application of ABA increases the number of lateral roots in the ABA receptor mutants (pyl8 and pyl9) in Arabidopsis, thus conferring to the plants a tolerance to salt [227]. These results indicate that the activation of ABA signaling displays enhanced salinity tolerance in crops and is triggered by an external application of ABA.
In wheat, ABA reduces salt stress damage by regulating proline content [228], and also by reducing the ROS levels in salt-grown plants [229]. Seed priming with ABA decreases Na+ content and increased K+ content in flag leaves, leading to increased number of grains per spike and grain yield of wheat under saline conditions, with up to 49% yield increase being reported [230]. In rice, exogenous application of 100 μM ABA solutions has improved plant performance by increasing OsP5CS1 and OsP5CR gene expression, which triggered proline accumulation, although the effect was varietal-dependent [231]. In wheat, the spraying of a moderate concentration of ABA (50 μmol·L−1) improved salt tolerance, while a higher concentration (100 μmol·L−1) had no significant impact [232]. Sorghum leaves were fed with an exogenous ABA to control shoot Na+ concentration and improve plant growth. The growth enhancement and lower Na+ content in shoots occur at a lower ABA concentration (≈10 mmol m−3) than a higher ABA concentration (≈40 mmol m−3 or above) under 150 mol m−3 NaCl treatment. A higher dose of ABA is needed to adapt plants to treatment with a lethal dose of NaCl (300 mol m−3). It is known that ABA acts by inducing transitional stomata closure to reduce transpiration and increase water use efficiency, thus lowering transpiration; this probably plays a significant role in reducing transporting Na+ from root to shoot [233]. In potatoes, exogenous application of ABA improves stomatal conductance and leaf relative water content under saline conditions, but the effect varies with genotype and method of ABA application, suggesting external ABA control over stomata functioning and better water saving under saline conditions [234].
Wheat LEA (late embryogenesis abundant) protein DHN-5, induced by salt and abscisic acid, can confer salt and osmotic stress tolerance, while Dhn-5 transgenic plants exhibit higher germination rates and leaf area, as well as better growth. The above salinity tolerance of the transgenic plant could be due to higher K+ accumulation in leaves and osmotic adjustment developed by active accumulation of proline [235]. In another study, a new member of the CIPK (calcineurin B-like protein-interacting protein kinase) gene family (TaCIPK29) has been identified in wheat. The TaCIPK29 transcription level increased after the treatment of ABA and NaCl. TaCIPK29 transgenic plant shows higher K+/Na+ ratio and increased activity of peroxidase (POD) and catalase (CAT) under salt stress [236]. A novel, ABA-inducible TaSC gene was cloned from a salt-tolerant wheat mutant, RH8706-49 [237], that operated in a CDPK pathway, enhancing intercellular K+/Na+ ratio and chloroplast function. Exogenous ABA treatment also promoted early salt stress-responding genes WESR1 and WESR2 in wheat [238]. A basic helix-loop-helix wheat gene (TabHLH1) mediates plant adaptation to osmotic stresses. This gene is associated with promoting stomata closure and increasing biomass production under salt and ABA treatment. The overexpression of TabHLH1 enhances leaf water retention capacity in transgenic tobacco, indicating better water saving to adapt under saline conditions [239].
While exogenous modulation of ABA levels in plants comes with improved water use efficiency (WUE), the resultant stomata closure may compromise the plant’s ability to assimilate CO2 (hence, biomass gain). This calls into question the long-term efficacy of such approaches for field-grown crops exposed to salinity.

3.8. Jasmonates

Another important set of PGRs are the jasmonates. Jasmonic acid (JA) and its methyl ester (MeJ) are known as jasmonates, and control a wide range of plant growth and developmental activities, as well as adaptive plant responses to a range of biotic and abiotic stressors, including salinity [242,243]. The biological activities of JA are significantly increased when plants are exposed to excess levels of salt [244], with stronger responses from salt-tolerant cultivars. This has prompted a suggestion to use jasmonic acid content as a proxy for salinity tolerance in plants [245,246,247]. Activation of the JA signaling pathway increases the accumulation of JA and increases plant salinity tolerance [248]. Consistent with this, the JA receptor mutant is associated with greater cell elongation under saline conditions to confer salinity tolerance [249]. Exogenous application of JA alleviates the toxic effects of salt by maintaining ion homeostasis, increasing ROS scavenging enzymatic activities, and improving stomatal functioning.
In wheat, exogenous 2 mM JA treatment alleviated salt stress by enhancing the activities and transcript levels of antioxidant enzymes, such as CAT, SOD, and APX. It also boosted the content of reduced glutathione (GSH) and carotenoids, thus decreasing the peroxidation of lipids [57]. Foliar JA sprays are beneficial for improving the grain yield of salt-grown soybeans [250]. The effectiveness of JA depends on its concentration and the level of salinity in the growing media [251].
JA induces biological and non-biological stress responses through the jasmonate signaling pathway [252]. For example, a salinity-responsive bread wheat gene TaAOC1 was constitutively expressed in both bread wheat and Arabidopsis, and was upregulated by exogenously supplied JA and ABA. The expression of TaAOC1 in both Arabidopsis and wheat restricted root growth, but enhanced salt tolerance and JA content by increasing SOD activity, indicating that JA was involved in the orchestration of salt stress response and developmental processes [248]. Other studies have shown that TaAOC1 and TaOPR1 [16,253] are the two genes that provide salt tolerance via both JA- and ABA-dependent pathways to promote expression of MYC2, a crucial component of abiotic stress response-signaling pathway [254]. Large-scale transcriptomic studies have shown that some JA-biosynthesis genes (e.g., AOC1, AOC2, AOS, LOX3 and OPR3) are up-regulated in roots under salt stress [255,256,257,258]. These findings suggest that JA signaling pathway is activated by salt stress and triggers an array of physiological and growth changes in plants. The TIFY gene family is regulated by salt and JA treatment, and transgenic lines over-expressing TdTIFY11a showed higher germination and growth rates under high-salinity conditions, indicating that it acts as jasmonic acid signaling [259]. A salt-responsive wheat gene TaEXPB23, associated with enhanced water retention ability and decreased osmotic potential, was upregulated by JA in transgenic tobacco [45].

3.9. Ethylene

Ethylene is a gaseous signaling molecule known as a stress-responsive hormone in plants [52,260]. It regulates a broad array of physiological and developmental responses [261,262] by cross-talking with other signaling molecules [263,264]. The functional effectiveness of ethylene depends on the sensitivity of plants to the hormone and its concentration in the cell [265,266,267,268]. Salt stress conditions cause a rapid increase of ethylene and its direct precursor ACC (1-aminocyclopropane-1-carboxylic acid) production inside the cell [269,270]. Higher ethylene production results in salt-sensitive phenotypes in many plants, such as rice, Arabidopsis, pepper, lettuce, spinach, and beetroot [269,271,272]. However, other studies have shown that the overproduction of endogenous ethylene or exogenous treatment of ethylene-releasing compounds, such as ethephon or ethylene precursors like ACC, may lead to increased salinity stress [273,274,275,276]. These results suggest that the effectiveness of ethylene in mitigating plant responses to salinity is crop- and genotypic-specific, and is controlled by the concentration of ethylene in a specific cellular compartment.
Similar to other PGRs, ethylene modulates salinity tolerance by maintaining tissue Na+/K+ homeostasis and inducing the antioxidant defense system [277,278,279]. Consistent with this, exogenous application of an ethylene-releasing compound significantly improves salinity tolerance in Arabidopsis. Exogenous 10 µM ACC (1-aminocyclopropane-1-carboxylic acid), an ethylene precursor, suppressed K+ loss and enhanced Na+ extrusion from the root, thus maintaining K+/Na+ homeostasis during short-term NaCl treatment in Arabidopsis [280]. While exogenous 30 μM ethephon (another ethylene-releasing agent) confers salt stress tolerance by increasing K+ ion content in shoots and roots rather than decreasing Na+ content in Arabidopsis, it also recovers salt-induced reductions in root growth [273]. In wheat, seed germination, as well as root and shoot length are significantly improved by different concentrations of ethephon under 100 mM NaCl treatment [281].
Ethylene response factors (ERFs) are key regulators in abiotic stress tolerance, including salinity. The transcription of wheat ERF gene (TaERF1) is induced by salinity, and overexpression of this gene activates stress-related genes that eventually increase salt stress tolerance in transgenic plants [282]. The seedlings of the TaERF3-overexpressing transgenic lines exhibit significantly enhanced tolerance to both drought and salt stresses compared to untransformed wheat [283]. Overexpression of an ethylene-responsive transcription factor (TdSHN1) from durum wheat resulted in the development of a thicker cuticle and lower stomatal density, thus reducing water loss in transgenic tobacco [284]. Transcripts of the lipid transfer protein gene (TaLTP1) were increased by salt and ethephon treatment in wheat [285]. LTPs enhance cell membrane integrity and ROS scavenging in transgenic potatoes [286]. They also result in reduced Na+ accumulation in transgenic tobacco [287]. A wheat aquaporin gene TaAQP8 conferred salt stress tolerance in transgenic tobacco by increasing the K+/Na+ ratio and Ca2+ content, and by reducing membrane damage and H2O2 accumulation [288]. Its transcript levels were induced by both ethylene and NaCl. Ethylene also induces many early response genes that are essential for ribosomal protein activation, chaperoning synthesis, ROS scavenging, and carbohydrate metabolites pathway [289].

4. Summary and Recommendations

When exposed to saline stress, plants display retarded growth and development and yield losses, and employ a range of mechanisms to deal with various constraints imposed by saline soils. Plant hormones play an important role in this process. Using exogenously applied PGRs remains a highly attractive option to plant growers, as a cost-effective method to induce salt tolerance genes and assist plants in adapting to hostile salinity conditions. However, the effectiveness of PGRs depends on the level of salt stress, genotype, timing, and methods of applications, as well as PGR concentrations. The issue is also complicated by the facts that plant hormones are involved in numerous developmental and adaptive responses (not only those related to salinity), and hormonal signaling pathways have a very significant overlap. Thus, elevation in the basal level of one of the PGRs could result in a major disturbance to some other signaling pathways, with pleiotropic effects for growth, development, and adaptation. This is specifically true for PGRs that modulate endogenous ROS and NO levels. The practical applicability of PGRs should also be considered in technological and economic contexts. Root treatment with PGRs reported in many papers is appropriate for laboratory-based studies, but has no place in the field. The aerial PGR sprays are more practical, but require significant technological developments (e.g., the use of surface surfactants, timing of spray application, etc.). The overall effects of PGR sprays will also be strongly dependent on environmental conditions (temperature, humidity, time of the day), as their penetration into the leaf will be largely determined by the extent of the stomata opening. The cost–benefit analysis of the efficacy of PGR ariel sprays should also be taken into account. We would like to illustrate the latter point by one simple example. In Yusuf et al. [177], the authors reported an 18% to 35% increase in seed yield in salt-grown peas, using a foliar spray of 0.125 mg L−1 of 24-epibrassinolide (EBL). The current cost of 10 mg of EBL from Sigma-Aldrich (Sigma-Aldrich Pty Ltd., NSW, Australia) is $588, and will be sufficient to make only 80 L of solution. The typical field rate of aerial spray application is 450 L ha−1 [290], so the cost of spraying of 1 ha will be about $3300 (EBL only). At the same time, the “target benchmark” for pea production in Australia is 8 tons ha−1 [291], with the commodity price being around $1000 per tonne in 2019 [292]. Thus, even a 30% increase in yield following EBL application will only result in a benefit of $2400 ha−1, which is clearly not enough to cover the cost of EBL application. The same logic is applicable to all other PGRs. Thus, all above beneficial reports of PGR application need to be taken with a “pinch of skepticism” and critically evaluated for their economic rationale.
In this context, we believe that future progress in the field may be achieved not by exogenous application of PGRs, but rather by understanding a causal link between PGRs and their downstream effectors mediating plants’ adaptation to salinity, and then incorporating these findings into a variety of plants via molecular breeding. This task, however, remains a great challenge, and can be only resolved by moving from whole-plant studies (employed by 95% of published papers) to more in-depth studies at the cellular level, using a modern range of biophysical and imaging techniques that allow quantification of the operation of key transport systems conferring plant ionic and oxidative homeostasis under stress conditions.

Author Contributions

Writing—original draft preparation, M.Q. and S.M.N.M.; writing—critical revision, M.Z. and S.S.; data analysis and editing, M.Z. and S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Grains Research and Development Corporation (GRDC) of Australia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Zhao, C.; Zhang, H.; Song, C.; Zhu, J.-K.; Shabala, S. Mechanisms of plant responses and adaptation to soil salinity. Innovation 2020, 1, 100017. [Google Scholar]
  2. Rengasamy, P. World salinization with emphasis on Australia. J. Exp. Bot. 2006, 57, 1017–1023. [Google Scholar] [PubMed] [Green Version]
  3. Hossain, M.S. Present scenario of global salt affected soils, its management and importance of salinity research. Int. Res. J. Biol. Sci. 2019, 1, 1–3. [Google Scholar]
  4. Shabala, S.; Bose, J.; Hedrich, R. Salt bladders: Do they matter? Trends Plant Sci 2014, 19, 687–691. [Google Scholar] [PubMed]
  5. Munns, R. Comparative physiology of salt and water stress. Plant Cell Environ. 2002, 25, 239–250. [Google Scholar]
  6. Wu, H.; Zhang, X.; Giraldo, J.P.; Shabala, S. It is not all about sodium: Revealing tissue specificity and signalling roles of potassium in plant responses to salt stress. Plant Soil 2018, 431, 1–17. [Google Scholar]
  7. Shabala, L.; Zhang, J.; Pottosin, I.; Bose, J.; Zhu, M.; Fuglsang, A.T.; Velarde-Buendia, A.; Massart, A.; Hill, C.B.; Roessner, U.; et al. Cell-type-specific H+-ATPase activity in root tissues enables K+ retention and mediates acclimation of barley (Hordeum vulgare) to salinity stress. Plant Physiol. 2016, 172, 2445–2458. [Google Scholar]
  8. Shabala, S. Salinity and programmed cell death: Unravelling mechanisms for ion specific signalling. J. Exp. Bot. 2009, 60, 709–712. [Google Scholar]
  9. Rubio, F.; Nieves-Cordones, M.; Horie, T.; Shabala, S. Doing ‘business as usual’comes with a cost: Evaluating energy cost of maintaining plant intracellular K+ homeostasis under saline conditions. N. Phytol. 2020, 225, 1097–1104. [Google Scholar]
  10. Bazihizina, N.; Colmer, T.D.; Cuin, T.A.; Mancuso, S.; Shabala, S. Friend or foe? Chloride patterning in halophytes. Trends Plant Sci. 2019, 24, 142–151. [Google Scholar]
  11. Teakle, N.L.; Tyerman, S.D. Mechanisms of Cl-transport contributing to salt tolerance. Plant Cell Environ. 2010, 33, 566–589. [Google Scholar] [PubMed]
  12. Sharma, P.; Jha, A.B.; Dubey, R.S.; Pessarakli, M. Reactive oxygen species, oxidative damage, and antioxidative defense mechanism in plants under stressful conditions. J. Bot. 2012. [Google Scholar] [CrossRef] [Green Version]
  13. Yang, Y.; Guo, Y. Unraveling salt stress signaling in plants. J. Integr. Plant Biol. 2018, 60, 796–804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Miller, G.; Suzuki, N.; Ciftci-Yilmaz, S.; Mittler, R. Reactive oxygen species homeostasis and signalling during drought and salinity stresses. Plant Cell Environ. 2010, 33, 453–467. [Google Scholar]
  15. Arya, A.; Nyamathulla, S.; Noordin, M.I.; Mohd, M.A. Antioxidant and hypoglycemic activities of leaf extracts of three popular Terminalia species. E-J. Chem. 2012, 9, 883–892. [Google Scholar] [CrossRef] [Green Version]
  16. Golldack, D.; Li, C.; Mohan, H.; Probst, N. Tolerance to drought and salt stress in plants: Unraveling the signaling networks. Front. Plant Sci. 2014, 5, 151. [Google Scholar]
  17. Genisel, M.; Erdal, S.; Kizilkaya, M. The mitigating effect of cysteine on growth inhibition in salt-stressed barley seeds is related to its own reducing capacity rather than its effects on antioxidant system. Plant Growth Regul. 2015, 75, 187–197. [Google Scholar] [CrossRef]
  18. Wang, W.-B.; Kim, Y.-H.; Lee, H.-S.; Kim, K.-Y.; Deng, X.-P.; Kwak, S.-S. Analysis of antioxidant enzyme activity during germination of alfalfa under salt and drought stresses. Plant Physiol. Biochem. 2009, 47, 570–577. [Google Scholar] [CrossRef]
  19. Pottosin, I.; Dobrovinskaya, O. Non-selective cation channels in plasma and vacuolar membranes and their contribution to K+ transport. J. Plant Physiol. 2014, 171, 732–742. [Google Scholar] [CrossRef]
  20. Demidchik, V.; Shabala, S.; Isayenkov, S.; Cuin, T.A.; Pottosin, I. Calcium transport across plant membranes: Mechanisms and functions. N. Phytol. 2018, 220, 49–69. [Google Scholar] [CrossRef] [Green Version]
  21. Demidchik, V.; Shabala, S. Mechanisms of cytosolic calcium elevation in plants: The role of ion channels, calcium extrusion systems and NADPH oxidase-mediated ‘ROS-Ca2+ Hub’. Funct. Plant Biol. 2018, 45, 9–27. [Google Scholar] [CrossRef]
  22. Zaman, B.; Niazi, B.; Athar, M.; Ahmad, M. Response of wheat plants to sodium and calcium ion interaction under saline environment. Int. J. Environ. Sci. Technol. 2005, 2, 7–12. [Google Scholar] [CrossRef] [Green Version]
  23. Dang, Y.; Routley, R.; McDonald, M.; Dalal, R.; Singh, D.; Orange, D.; Mann, M. Subsoil constraints in Vertosols: Crop water use, nutrient concentration, and grain yields of bread wheat, durum wheat, barley, chickpea, and canola. Aust. J. Agric. Res. 2006, 57, 983–998. [Google Scholar] [CrossRef] [Green Version]
  24. McDonald, G.K.; Taylor, J.; Verbyla, A.; Kuchel, H. Assessing the importance of subsoil constraints to yield of wheat and its implications for yield improvement. Crop Pasture Sci. 2013, 63, 1043–1065. [Google Scholar] [CrossRef]
  25. Shabala, S.; Munns, R. Salinity stress: Physiological constraints and adaptive mechanisms. Plant Stress Physiol. 2012, 1, 59–93. [Google Scholar]
  26. Orton, T.G.; Mallawaarachchi, T.; Pringle, M.J.; Menzies, N.W.; Dalal, R.C.; Kopittke, P.M.; Searle, R.; Hochman, Z.; Dang, Y.P. Quantifying the economic impact of soil constraints on Australian agriculture: A case-study of wheat. Land Degrad. Dev. 2018, 29, 3866–3875. [Google Scholar] [CrossRef]
  27. Ashraf, M.; Athar, H.; Harris, P.; Kwon, T. Some prospective strategies for improving crop salt tolerance. Adv. Agron. 2008, 97, 45–110. [Google Scholar]
  28. Melero, S.; Madejón, E.; Ruiz, J.C.; Herencia, J.F. Chemical and biochemical properties of a clay soil under dryland agriculture system as affected by organic fertilization. Eur. J. Agron. 2007, 26, 327–334. [Google Scholar] [CrossRef] [Green Version]
  29. Shahbaz, M.; Ashraf, M. Improving salinity tolerance in cereals. Crit. Rev. Plant Sci. 2013, 32, 237–249. [Google Scholar]
  30. Zahir, Z.A.; Akhtar, S.S.; Ahmad, M.; Nadeem, S.M.; Nadeem, S.M. Comparative effectiveness of Enterobacter aerogenes and Pseudomonas fluorescens for mitigating the depressing effect of brackish water on maize. Int. J. Agric. Biol. 2012, 14, 337–344. [Google Scholar]
  31. Akhtar, S.S.; Andersen, M.N.; Naveed, M.; Zahir, Z.A.; Liu, F. Interactive effect of biochar and plant growth-promoting bacterial endophytes on ameliorating salinity stress in maize. Funct. Plant Biol. 2015, 42, 770–781. [Google Scholar] [CrossRef]
  32. Hanay, A.; Büyüksönmez, F.; Kiziloglu, F.M.; Canbolat, M.Y. Reclamation of saline-sodic soils with gypsum and MSW compost. Compos. Sci. Util. 2004, 12, 175–179. [Google Scholar] [CrossRef]
  33. Akhtar, S.S.; Andersen, M.N.; Liu, F. Residual effects of biochar on improving growth, physiology and yield of wheat under salt stress. Agric. Water Manag. 2015, 158, 61–68. [Google Scholar] [CrossRef]
  34. Fahad, S.; Hussain, S.; Matloob, A.; Khan, F.A.; Khaliq, A.; Saud, S.; Hassan, S.; Shan, D.; Khan, F.; Ullah, N.; et al. Phytohormones and plant responses to salinity stress: A review. Plant Growth Regul. 2015, 75, 391–404. [Google Scholar] [CrossRef]
  35. Flowers, T. Improving crop salt tolerance. J. Exp. Bot. 2004, 55, 307–319. [Google Scholar] [CrossRef] [PubMed]
  36. Chinnusamy, V.; Jagendorf, A.; Zhu, J.K. Understanding and improving salt tolerance in plants. Crop Sci. 2005, 45, 437–448. [Google Scholar] [CrossRef] [Green Version]
  37. Liang, W.; Ma, X.; Wan, P.; Liu, L. Plant salt-tolerance mechanism: A review. Biochem. Biophys. Res. Commun. 2018, 495, 286–291. [Google Scholar] [CrossRef]
  38. Chan, R.; Trucco, F.; Otegui, M. Why are second-generation transgenic crops not yet available in the market? J. Exp. Bot. 2020, 71, 6876–6880. [Google Scholar] [CrossRef] [PubMed]
  39. Ricroch, A.E.; Guillaume-Hofnung, M.; Kuntz, M. The ethical concerns about transgenic crops. Biochem. J. 2018, 475, 803–811. [Google Scholar] [CrossRef] [PubMed]
  40. Shew, A.M.; Nalley, L.L.; Snell, H.A.; Nayga, R.M., Jr.; Dixon, B.L. CRISPR versus GMOs: Public acceptance and valuation. Glob. Food Secur. 2018, 19, 71–80. [Google Scholar] [CrossRef]
  41. Rademacher, W. Plant growth regulators: Backgrounds and uses in plant production. J. Plant Growth Regul. 2015, 34, 845–872. [Google Scholar] [CrossRef]
  42. Khan, N.; Bano, A.; Zandi, P. Effects of exogenously applied plant growth regulators in combination with PGPR on the physiology and root growth of chickpea (Cicer arietinum) and their role in drought tolerance. J. Plant Interact. 2018, 13, 239–247. [Google Scholar] [CrossRef] [Green Version]
  43. Nadeem, S.M.; Ahmad, M.; Zahir, Z.A.; Javaid, A.; Ashraf, M. The role of mycorrhizae and plant growth promoting rhizobacteria (PGPR) in improving crop productivity under stressful environments. Biotechnol. Adv. 2014, 32, 429–448. [Google Scholar] [CrossRef] [PubMed]
  44. Keswani, C.; Prakash, O.; Bharti, N.; Vílchez, J.I.; Sansinenea, E.; Lally, R.D.; Borriss, R.; Singh, S.P.; Gupta, V.K.; Fraceto, L.F.; et al. Re-addressing the biosafety issues of plant growth promoting rhizobacteria. Sci. Total Environ. 2019, 690, 841–852. [Google Scholar] [CrossRef]
  45. Han, Y.Y.; Li, A.X.; Li, F.; Zhao, R.M.; Wang, W. Characterization of a wheat (Triticum aestivum L.) expansin gene, TaEXPB23, involved in the abiotic stress response and phytohormone regulation. Plant Physiol. Biochem. 2012, 54, 49–58. [Google Scholar] [CrossRef]
  46. Fishel, F.M. Plant Growth Regulators; University of Florida: Gainesville, FL, USA, 2006; p. PI-102. [Google Scholar]
  47. Hopkins, W.G. Introduction to Plant Physiology; John Wiley and Sons: Hoboken, NJ, USA, 1999. [Google Scholar]
  48. Avery, G.S.J. The Growth Hormones Found in Plants; 1937; Volume XXXVII, pp. 317–322. Available online: https://kb.osu.edu/bitstream/handle/1811/2894/V37N06_317.pdf?sequence=1 (accessed on 23 May 2021).
  49. Pál, M.; Janda, T.; Szalai, G. Interactions between plant hormones and thiol-related heavy metal chelators. Plant Growth Regul. 2018, 85, 173–185. [Google Scholar] [CrossRef] [Green Version]
  50. Bücker-Neto, L.; Paiva, A.L.S.; Machado, R.D.; Arenhart, R.A.; Margis-Pinheiro, M. Interactions between plant hormones and heavy metals responses. Genet. Mol. Biol. 2017, 40, 373–386. [Google Scholar] [CrossRef]
  51. Datta, K.; Varma, S.; Angrish, R.; Kumar, B.; Kumari, P. Alleviation of salt stress by plant growth regulators in Triticum aestivum L. Biol. Plant 1997, 40, 269–275. [Google Scholar] [CrossRef]
  52. Yu, Z.; Duan, X.; Luo, L.; Dai, S.; Ding, Z.; Xia, G. How plant hormones mediate salt stress responses. Trends Plant Sci. 2020. [Google Scholar] [CrossRef]
  53. Wani, S.H.; Kumar, V.; Shriram, V.; Sah, S.K. Phytohormones and their metabolic engineering for abiotic stress tolerance in crop plants. Crop J. 2016, 4, 162–176. [Google Scholar] [CrossRef] [Green Version]
  54. Iqbal, M.; Ashraf, M. Seed treatment with auxins modulates growth and ion partitioning in salt-stressed wheat plants. J. Integr. Plant Biol. 2007, 49, 1003–1015. [Google Scholar] [CrossRef]
  55. Faghih, S.; Zarei, A.; Ghobadi, C. Positive effects of plant growth regulators on physiology responses of Fragaria × ananassa cv.‘Camarosa’under salt stress. Int. J. Fruit Sci. 2019, 19, 104–114. [Google Scholar] [CrossRef]
  56. Rhaman, M.S.; Imran, S.; Rauf, F.; Khatun, M.; Baskin, C.C.; Murata, Y.; Hasanuzzaman, M. Seed Priming with Phytohormones: An Effective Approach for the Mitigation of Abiotic Stress. Plants 2021, 10, 37. [Google Scholar] [CrossRef] [PubMed]
  57. Qiu, Z.; Guo, J.; Zhu, A.; Zhang, L.; Zhang, M. Exogenous jasmonic acid can enhance tolerance of wheat seedlings to salt stress. Ecotoxicol. Environ. Saf. 2014, 104, 202–208. [Google Scholar] [CrossRef]
  58. Ashraf, M.; Karim, F.; Rasul, E. Interactive effects of gibberellic acid (GA 3) and salt stress on growth, ion accumulation and photosynthetic capacity of two spring wheat (Triticum aestivum L.) cultivars differing in salt tolerance. Plant Growth Regul. 2002, 36, 49–59. [Google Scholar] [CrossRef]
  59. Kaydan, D.; Yagmur, M.; Okut, N. Effects of salicylic acid on the growth and some physiological characters in salt stressed wheat (Triticum aestivum L.). Tarim Bilimleri Derg. 2007, 13, 114–119. [Google Scholar]
  60. Iqbal, M.; Ashraf, M.; Jamil, A. Seed enhancement with cytokinins: Changes in growth and grain yield in salt stressed wheat plants. Plant Growth Regul. 2006, 50, 29–39. [Google Scholar] [CrossRef]
  61. Iqbal, M.; Ashraf, M. Presowing seed treatment with cytokinins and its effect on growth, photosynthetic rate, ionic levels and yield of two wheat cultivars differing in salt tolerance. J. Integr. Plant Biol. 2005, 47, 1315–1325. [Google Scholar] [CrossRef]
  62. Iqbal, M.; Ashraf, M. Gibberellic acid mediated induction of salt tolerance in wheat plants: Growth, ionic partitioning, photosynthesis, yield and hormonal homeostasis. Environ. Exp. Bot. 2013, 86, 76–85. [Google Scholar] [CrossRef]
  63. Mittler, R.; Vanderauwera, S.; Suzuki, N.; Miller, G.; Tognetti, V.B.; Vandepoele, K.; Gollery, M.; Shulaev, V.; Breusegem, F.V. ROS signaling: The new wave? Trends Plant Sci. 2011, 16, 300–309. [Google Scholar] [CrossRef]
  64. Tognetti, V.B.; Mühlenbock, P.; Van Breusegem, F. Stress homeostasis–the redox and auxin perspective. Plant Cell Environ. 2012, 35, 321–333. [Google Scholar] [CrossRef] [PubMed]
  65. López, M.L.; Peralta-Videa, J.R.; Benitez, T.; Gardea-Torresdey, J.L. Enhancement of lead uptake by alfalfa (Medicago sativa) using EDTA and a plant growth promoter. Chemosphere 2005, 61, 595–598. [Google Scholar] [CrossRef] [PubMed]
  66. Hadi, F.; Bano, A.; Fuller, M.P. The improved phytoextraction of lead (Pb) and the growth of maize (Zea mays L.): The role of plant growth regulators (GA3 and IAA) and EDTA alone and in combinations. Chemosphere 2010, 80, 457–462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Fässler, E.; Evangelou, M.W.; Robinson, B.H.; Schulin, R. Effects of indole-3-acetic acid (IAA) on sunflower growth and heavy metal uptake in combination with ethylene diamine disuccinic acid (EDDS). Chemosphere 2010, 80, 901–907. [Google Scholar] [CrossRef]
  68. Davies, P.J. Regulatory Factors in Hormone Action: Level, Location and Signal Transduction. In Plant Hormones; Springer: Dordrecht, The Netherlands, 2010; pp. 16–35. [Google Scholar]
  69. Bennett, T.; Hines, G.; van Rongen, M.; Waldie, T.; Sawchuk, M.G.; Scarpella, E.; Ljung, K. Connective auxin transport in the shoot facilitates communication between shoot apices. PLoS Biol. 2016, 14, e1002446. [Google Scholar] [CrossRef]
  70. Leyser, O. Auxin signaling. Plant Physiol. 2018, 176, 465–479. [Google Scholar] [CrossRef] [Green Version]
  71. Brumos, J.; Robles, L.M.; Yun, J.; Vu, T.C.; Jackson, S.; Alonso, J.M.; Stepanova, A.N. Local auxin biosynthesis is a key regulator of plant development. Dev. Cell 2018, 47, 306–318.e5. [Google Scholar] [CrossRef] [Green Version]
  72. Lau, S.; Jürgens, G.; De Smet, I. The evolving complexity of the auxin pathway. Plant Cell 2008, 20, 1738–1746. [Google Scholar] [CrossRef] [Green Version]
  73. Heisler, M.G.; Byrne, M.E. Progress in understanding the role of auxin in lateral organ development in plants. Curr. Opin. Plant Biol. 2020, 53, 73–79. [Google Scholar] [CrossRef]
  74. Perrot-Rechenmann, C. Cellular responses to auxin: Division versus expansion. Cold Spring Harb. Perspect. Biol. 2010, 2, a001446. [Google Scholar] [CrossRef]
  75. Majda, M.; Robert, S. The role of auxin in cell wall expansion. Int. J. Mol. Sci. 2018, 19, 951. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Hager, A. Role of the plasma membrane H+-ATPase in auxin-induced elongation growth: Historical and new aspects. J. Plant Res. 2003, 116, 483–505. [Google Scholar] [CrossRef]
  77. Du, M.; Spalding, E.P.; Gray, W.M. Rapid Auxin-Mediated Cell Expansion. Annu. Rev. Plant Biol. 2020, 71, 379–402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Julkowska, M.M.; Hoefsloot, H.C.; Mol, S.; Feron, R.; de Boer, G.-J.; Haring, M.A.; Testerink, C. Capturing Arabidopsis root architecture dynamics with ROOT-FIT reveals diversity in responses to salinity. Plant Physiol. 2014, 166, 1387–1402. [Google Scholar] [CrossRef] [Green Version]
  79. Munns, R. Genes and salt tolerance: Bringing them together. N. Phytol. 2005, 167, 645–663. [Google Scholar] [CrossRef] [PubMed]
  80. Fu, Y.; Yang, Y.; Chen, S.; Ning, N.; Hu, H. Arabidopsis IAR4 modulates primary root growth under salt stress through ROS-mediated modulation of auxin distribution. Front. Plant Sci. 2019, 10, 522. [Google Scholar] [CrossRef]
  81. Korver, R.A.; Koevoets, I.T.; Testerink, C. Out of shape during stress: A key role for auxin. Trends Plant Sci. 2018, 23, 783–793. [Google Scholar] [CrossRef]
  82. Liu, W.; Li, R.-J.; Han, T.-T.; Cai, W.; Fu, Z.-W.; Lu, Y.-T. Salt stress reduces root meristem size by nitric oxide-mediated modulation of auxin accumulation and signaling in Arabidopsis. Plant Physiol. 2015, 168, 343–356. [Google Scholar] [CrossRef] [Green Version]
  83. Jiang, K.; Moe-Lange, J.; Hennet, L.; Feldman, L.J. Salt stress affects the redox status of Arabidopsis root meristems. Front. Plant Sci. 2016, 7, 81. [Google Scholar] [CrossRef] [Green Version]
  84. Khalid, A.; Aftab, F. Effect of exogenous application of IAA and GA 3 on growth, protein content, and antioxidant enzymes of Solanum tuberosum L. grown in vitro under salt stress. Vitr. Cell. Dev. Biol. Plant 2020, 56, 1–13. [Google Scholar]
  85. Ivanov, V.; Bystrova, E.; Seregin, I. Comparative impacts of heavy metals on root growth as related to their specificity and selectivity. Russ. J. Plant Physiol. 2003, 50, 398–406. [Google Scholar] [CrossRef]
  86. Sun, F.; Zhang, W.; Hu, H.; Li, B.; Wang, Y.; Zhao, Y.; Li, K.; Liu, M.; Li, X. Salt modulates gravity signaling pathway to regulate growth direction of primary roots in Arabidopsis. Plant Physiol. 2008, 146, 178–188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Prasad, M.N.V. Heavy Metal Stress in Plants: From Biomolecules to Ecosystems; Springer Science & Business Media: Berlin, Germany, 2013. [Google Scholar]
  88. Petersson, S.V.; Johansson, A.I.; Kowalczyk, M.; Makoveychuk, A.; Wang, J.Y.; Moritz, T.; Grebe, M.; Benfey, P.N.; Sandberg, G.; Ljung, K. An auxin gradient and maximum in the Arabidopsis root apex shown by high-resolution cell-specific analysis of IAA distribution and synthesis. Plant Cell 2009, 21, 1659–1668. [Google Scholar] [CrossRef]
  89. Wang, Y.; Li, K.; Li, X. Auxin redistribution modulates plastic development of root system architecture under salt stress in Arabidopsis thaliana. J. Plant Physiol. 2009, 166, 1637–1645. [Google Scholar] [CrossRef] [PubMed]
  90. He, F.; Xu, C.; Fu, X.; Shen, Y.; Guo, L.; Leng, M.; Luo, K. The MicroRNA390/TRANS-acting short interfering RNA3 module mediates lateral root growth under salt stress via the auxin pathway. Plant Physiol. 2018, 177, 775–791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Doorn, M.M.V.; Merl-Pham, J.; Ghirardo, A.; Fink, S.; Polle, A.; Schnitzler, J.P.; Rosenkranz, M. Root isoprene formation alters lateral root development. Plant Cell Environ. 2020, 43, 2207–2223. [Google Scholar] [CrossRef] [PubMed]
  92. Simon, L.; Bousquet, J.; Lévesque, R.C.; Lalonde, M. Origin and diversification of endomycorrhizal fungi and coincidence with vascular land plants. Nature 1993, 363, 67–69. [Google Scholar] [CrossRef]
  93. Aoi, Y.; Tanaka, K.; Cook, S.D.; Hayashi, K.-I.; Kasahara, H. GH3 auxin-amido synthetases alter the ratio of indole-3-acetic acid and phenylacetic acid in Arabidopsis. Plant Cell Physiol. 2020, 61, 596–605. [Google Scholar] [CrossRef]
  94. Kazan, K. Auxin and the integration of environmental signals into plant root development. Ann. Bot. 2013, 112, 1655–1665. [Google Scholar] [CrossRef] [Green Version]
  95. Zörb, C.; Geilfus, C.-M.; Mühling, K.H.; Ludwig-Müller, J. The influence of salt stress on ABA and auxin concentrations in two maize cultivars differing in salt resistance. J. Plant Physiol. 2013, 170, 220–224. [Google Scholar] [CrossRef]
  96. Ji, J.; Yuan, D.; Jin, C.; Wang, G.; Li, X.; Guan, C. Enhancement of growth and salt tolerance of rice seedlings (Oryza sativa L.) by regulating ethylene production with a novel halotolerant PGPR strain Glutamicibacter sp. YD01 containing ACC deaminase activity. Acta Physiol. Plant 2020, 42, 1–17. [Google Scholar] [CrossRef]
  97. Zhang, S.; Gan, Y.; Xu, B. Mechanisms of the IAA and ACC-deaminase producing strain of Trichoderma longibrachiatum T6 in enhancing wheat seedling tolerance to NaCl stress. BMC Plant Biol. 2019, 19, 22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Kaya, C.; Ashraf, M.; Dikilitas, M.; Tuna, A.L. Alleviation of salt stress-induced adverse effects on maize plants by exogenous application of indoleacetic acid (IAA) and inorganic nutrients—A field trial. Aust. J. Crop Sci. 2013, 7, 249. [Google Scholar]
  99. Javid, M.G.; Sorooshzadeh, A.; Sanavy, S.A.M.M.; Allahdadi, I.; Moradi, F. Effects of the exogenous application of auxin and cytokinin on carbohydrate accumulation in grains of rice under salt stress. Plant Growth Regul. 2011, 65, 305–313. [Google Scholar] [CrossRef]
  100. Iqbal, M.; Ashraf, M. Alleviation of salinity-induced perturbations in ionic and hormonal concentrations in spring wheat through seed preconditioning in synthetic auxins. Acta Physiol. Plant 2013, 35, 1093–1112. [Google Scholar] [CrossRef]
  101. Nawaz, A.; Shahbaz, M.; Asadullah; Imran, A.; Marghoob, M.U.; Imtiaz, M.; Mubeen, F. Potential of salt tolerant PGPR in growth and yield augmentation of wheat (Triticum aestivum L.) under saline conditions. Front. Microbiol. 2020, 11, 2019. [Google Scholar] [CrossRef]
  102. Guo, Y.; Jiang, Q.; Hu, Z.; Sun, X.; Fan, S.; Zhang, H. Function of the auxin-responsive gene TaSAUR75 under salt and drought stress. Crop J. 2018, 6, 181–190. [Google Scholar] [CrossRef]
  103. Talboys, P.J.; Healey, J.R.; Withers, P.J.; Jones, D.L. Phosphate depletion modulates auxin transport in Triticum aestivum leading to altered root branching. J. Exp. Bot. 2014, 65, 5023–5032. [Google Scholar] [CrossRef] [Green Version]
  104. Aina, O.; Quesenberry, K.; Gallo, M. Thidiazuron-induced tissue culture regeneration from quartered-seed explants of Arachis paraguariensis. Crop Sci. 2012, 52, 1076–1083. [Google Scholar]
  105. Mok, D.W.; Mok, M.C. Cytokinin metabolism and action. Annu. Rev. Plant Biol. 2001, 52, 89–118. [Google Scholar] [CrossRef]
  106. Sakakibara, H. Cytokinins: Activity, biosynthesis, and translocation. Annu. Rev. Plant Biol. 2006, 57, 431–449. [Google Scholar] [CrossRef] [Green Version]
  107. Stirk, W.A.; Gold, J.D.; Novák, O.; Strnad, M.; Van Staden, J. Changes in endogenous cytokinins during germination and seedling establishment of Tagetes minuta L. Plant Growth Regul. 2005, 47, 1–7. [Google Scholar] [CrossRef]
  108. Bishopp, A.; Lehesranta, S.; Vatén, A.; Help, H.; El-Showk, S.; Scheres, B.; Helariutta, K.; Mähönen, A.P.; Sakakibara, H.; Helariutta, Y.H. Phloem-transported cytokinin regulates polar auxin transport and maintains vascular pattern in the root meristem. Curr. Biol. 2011, 21, 927–932. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Kiba, T.; Takei, K.; Kojima, M.; Sakakibara, H. Side-chain modification of cytokinins controls shoot growth in Arabidopsis. Dev. Cell 2013, 27, 452–461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Li, S.-M.; Zheng, H.-X.; Zhang, X.-S.; Sui, N. Cytokinins as central regulators during plant growth and stress response. Plant Cell Rep. 2021, 40, 271–282. [Google Scholar] [CrossRef] [PubMed]
  111. Hyoung, S.; Cho, S.H.; Chung, J.H.; So, W.M.; Cui, M.H.; Shin, J.S. Cytokinin oxidase PpCKX1 plays regulatory roles in development and enhances dehydration and salt tolerance in Physcomitrella patens. Plant Cell Rep. 2020, 39, 419–430. [Google Scholar] [CrossRef] [PubMed]
  112. Nishiyama, R.; Watanabe, Y.; Fujita, Y.; Le, D.T.; Kojima, M.; Werner, T.; Vankova, R.; Yamaguchi-Shinozaki, K.; Shinozaki, K.; Kakimoto, T.; et al. Analysis of cytokinin mutants and regulation of cytokinin metabolic genes reveals important regulatory roles of cytokinins in drought, salt and abscisic acid responses, and abscisic acid biosynthesis. Plant Cell 2011, 23, 2169–2183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Wang, Y.; Shen, W.; Chan, Z.; Wu, Y. Endogenous cytokinin overproduction modulates ROS homeostasis and decreases salt stress resistance in Arabidopsis thaliana. Front. Plant Sci. 2015, 6, 1004. [Google Scholar] [CrossRef] [Green Version]
  114. Cortleven, A.; Leuendorf, J.E.; Frank, M.; Pezzetta, D.; Bolt, S.; Schmülling, T. Cytokinin action in response to abiotic and biotic stresses in plants. Plant Cell Environ. 2019, 42, 998–1018. [Google Scholar] [CrossRef]
  115. Tran, L.-S.P.; Urao, T.; Qin, F.; Maruyama, K.; Kakimoto, T.; Shinozaki, K.; Yamaguchi-Shinozaki, K. Functional analysis of AHK1/ATHK1 and cytokinin receptor histidine kinases in response to abscisic acid, drought, and salt stress in Arabidopsis. Proc. Natl. Acad. Sci. USA 2007, 104, 20623–20628. [Google Scholar] [CrossRef] [Green Version]
  116. Nishiyama, R.; Le, D.T.; Watanabe, Y.; Matsui, A.; Tanaka, M.; Seki, M.; Yamaguchi-Shinozaki, K.; Shinozaki, K.; Tran, L.P. Transcriptome analyses of a salt-tolerant cytokinin-deficient mutant reveal differential regulation of salt stress response by cytokinin deficiency. PLoS ONE 2012, 7, e32124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Aremu, A.O.; Masondo, N.A.; Sunmonu, T.O.; Kulkarni, M.G.; Zatloukal, M.; Spichal, L.; Doležal, K.; Staden, J.V. A novel inhibitor of cytokinin degradation (INCYDE) influences the biochemical parameters and photosynthetic apparatus in NaCl-stressed tomato plants. Planta 2014, 240, 877–889. [Google Scholar] [CrossRef] [PubMed]
  118. Joshi, R.; Sahoo, K.K.; Tripathi, A.K.; Kumar, R.; Gupta, B.K.; Pareek, A.; Singla-Pareek, S.L. Knockdown of an inflorescence meristem-specific cytokinin oxidase–OsCKX2 in rice reduces yield penalty under salinity stress condition. Plant Cell Environ. 2018, 41, 936–946. [Google Scholar] [CrossRef] [PubMed]
  119. Ashikari, M.; Sakakibara, H.; Lin, S.; Yamamoto, T.; Takashi, T.; Nishimura, A.; Angeles, E.R.; Qian, Q.; Kitano, H.; Matsuoka, M. Cytokinin oxidase regulates rice grain production. Science 2005, 309, 741–745. [Google Scholar] [CrossRef]
  120. Yin, W.; Xiao, Y.; Niu, M.; Meng, W.; Li, L.; Zhang, X.; Liu, D.; Zhang, G.; Qian, Y.; Sun, Z.; et al. ARGONAUTE2 enhances grain length and salt tolerance by activating BIG GRAIN3 to modulate cytokinin distribution in rice. Plant Cell 2020, 32, 2292–2306. [Google Scholar] [CrossRef]
  121. Gadallah, M. Effects of kinetin on growth, grain yield and some mineral elements in wheat plants growing under excess salinity and oxygen deficiency. Plant Growth Regul. 1999, 27, 63–74. [Google Scholar] [CrossRef]
  122. Osugi, A.; Sakakibara, H. Q&A: How do plants respond to cytokinins and what is their importance? BMC Biol. 2015, 13, 102. [Google Scholar]
  123. Kakimoto, T. CKI1, a histidine kinase homolog implicated in cytokinin signal transduction. Science 1996, 274, 982–985. [Google Scholar] [CrossRef]
  124. Zwack, P.J.; De Clercq, I.; Howton, T.C.; Hallmark, H.T.; Hurny, A.; Keshishian, E.A.; Parish, A.M.; Benkova, E.; Mukhtar, M.S.; Breusegem, F.V.; et al. Cytokinin response factor 6 represses cytokinin-associated genes during oxidative stress. Plant Physiol. 2016, 172, 1249–1258. [Google Scholar] [CrossRef] [Green Version]
  125. Ma, X.; Feng, D.-S.; Wang, H.-G.; Li, X.-F.; Kong, L.-R. Cloning and expression analysis of wheat cytokinin oxidase/dehydrogenase gene TaCKX3. Plant Mol. Biol. Rep. 2011, 29, 98–105. [Google Scholar] [CrossRef]
  126. Jabłoński, B.; Ogonowska, H.; Szala, K.; Bajguz, A.; Orczyk, W.; Nadolska-Orczyk, A. Silencing of TaCKX1 mediates expression of other TaCKX genes to increase grain yield in wheat. Int. J. Mol. Sci. 2020, 21, 4809. [Google Scholar] [CrossRef] [PubMed]
  127. Mason, M.G.; Jha, D.; Salt, D.E.; Tester, M.; Hill, K.; Kieber, J.J.; Schaller, G.E. Type-B response regulators ARR1 and ARR12 regulate expression of AtHKT1; 1 and accumulation of sodium in Arabidopsis shoots. Plant J. 2010, 64, 753–763. [Google Scholar] [CrossRef] [PubMed]
  128. Rashotte, A.M.; Goertzen, L.R. The CRF domain defines cytokinin response factor proteins in plants. BMC Plant Biol. 2010, 10, 74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Rashotte, A.M.; Mason, M.G.; Hutchison, C.E.; Ferreira, F.J.; Schaller, G.E.; Kieber, J.J. A subset of Arabidopsis AP2 transcription factors mediates cytokinin responses in concert with a two-component pathway. Proc. Natl. Acad. Sci. USA 2006, 103, 11081–11085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Nishiyama, R.; Watanabe, Y.; Leyva-Gonzalez, M.A.; Van Ha, C.; Fujita, Y.; Tanaka, M.; Seki, M.; Yamaguchi-Shinozaki, K.; Shinozaki, K.; Herrera-Estrella, L.; et al. Arabidopsis AHP2, AHP3, and AHP5 histidine phosphotransfer proteins function as redundant negative regulators of drought stress response. Proc. Natl. Acad. Sci. USA 2013, 110, 4840–4845. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Mira-Rodado, V.; Veerabagu, M.; Witthöft, J.; Teply, J.; Harter, K.; Desikan, R. Identification of two-component system elements downstream of AHK5 in the stomatal closure response of Arabidopsis thaliana. Plant Signal. Behav. 2012, 7, 1467–1476. [Google Scholar] [CrossRef] [Green Version]
  132. Davenport, R.J.; Muñoz-Mayor, A.; Jha, D.; Essah, P.A.; Rus, A.; Tester, M. The Na+ transporter AtHKT1; 1 controls retrieval of Na+ from the xylem in Arabidopsis. Plant Cell Environ. 2007, 30, 497–507. [Google Scholar] [CrossRef]
  133. Nguyen, K.H.; Van Ha, C.; Nishiyama, R.; Watanabe, Y.; Leyva-González, M.A.; Fujita, Y.; Tran, U.T.; Li, W.; Tanaka, M.; Seki, M.; et al. Arabidopsis type B cytokinin response regulators ARR1, ARR10, and ARR12 negatively regulate plant responses to drought. Proc. Natl. Acad. Sci. USA 2016, 113, 3090–3095. [Google Scholar] [CrossRef] [Green Version]
  134. Li, S.; An, Y.; Hailati, S.; Zhang, J.; Cao, Y.; Liu, Y.; Geng, J.; Hu, T.; Yang, P. Overexpression of the cytokinin oxidase/dehydrogenase (CKX) from Medicago sativa enhanced salt stress tolerance of Arabidopsis. J. Plant Biol. 2019, 62, 374–386. [Google Scholar] [CrossRef]
  135. Corpas, F.J.; Palma, J.M.; Del Río, L.A.; Barroso, J.B. Evidence supporting the existence of L-arginine-dependent nitric oxide synthase activity in plants. N. Phytol. 2009, 184, 9–14. [Google Scholar] [CrossRef]
  136. Rockel, P.; Strube, F.; Rockel, A.; Wildt, J.; Kaiser, W.M. Regulation of nitric oxide (NO) production by plant nitrate reductase in vivo and in vitro. J. Exp. Bot. 2002, 53, 103–110. [Google Scholar] [CrossRef] [PubMed]
  137. Wilson, I.D.; Neill, S.J.; Hancock, J.T. Nitric oxide synthesis and signalling in plants. Plant Cell Environ. 2008, 31, 622–631. [Google Scholar] [CrossRef] [PubMed]
  138. Corpas, F.J.; Barroso, J.B.; Carreras, A.; Valderrama, R.; Palma, J.M.; León, A.M.; Sandalio, L.M.; Río, L.A.D. Constitutive arginine-dependent nitric oxide synthase activity in different organs of pea seedlings during plant development. Planta 2006, 224, 246–254. [Google Scholar] [CrossRef] [PubMed]
  139. Asgher, M.; Per, T.S.; Masood, A.; Fatma, M.; Freschi, L.; Corpas, F.J.; Khan, N.A. Nitric oxide signaling and its crosstalk with other plant growth regulators in plant responses to abiotic stress. Environ. Sci. Pollut. Res. 2017, 24, 2273–2285. [Google Scholar] [CrossRef]
  140. Sami, F.; Faizan, M.; Faraz, A.; Siddiqui, H.; Yusuf, M.; Hayat, S. Nitric oxide-mediated integrative alterations in plant metabolism to confer abiotic stress tolerance, NO crosstalk with phytohormones and NO-mediated post translational modifications in modulating diverse plant stress. Nitric Oxide 2018, 73, 22–38. [Google Scholar] [CrossRef] [PubMed]
  141. Khan, M.N.; Mobin, M.; Abbas, Z.K.; Siddiqui, M.H. Nitric oxide-induced synthesis of hydrogen sulfide alleviates osmotic stress in wheat seedlings through sustaining antioxidant enzymes, osmolyte accumulation and cysteine homeostasis. Nitric Oxide 2017, 68, 91–102. [Google Scholar] [CrossRef]
  142. Siddiqui, M.H.; Al-Whaibi, M.H.; Basalah, M.O. Role of nitric oxide in tolerance of plants to abiotic stress. Protoplasma 2011, 248, 447–455. [Google Scholar] [CrossRef] [PubMed]
  143. Khan, M.N.; Siddiqui, M.H.; Mohammad, F.; Naeem, M. Interactive role of nitric oxide and calcium chloride in enhancing tolerance to salt stress. Nitric Oxide 2012, 27, 210–218. [Google Scholar] [CrossRef]
  144. Shi, H.; Liu, W.; Wei, Y.; Ye, T. Integration of auxin/indole-3-acetic acid 17 and RGA-LIKE3 confers salt stress resistance through stabilization by nitric oxide in Arabidopsis. J. Exp. Bot. 2017, 68, 1239–1249. [Google Scholar] [CrossRef]
  145. Zheng, C.; Jiang, D.; Liu, F.; Dai, T.; Liu, W.; Jing, Q.; Cao, W. Exogenous nitric oxide improves seed germination in wheat against mitochondrial oxidative damage induced by high salinity. Environ. Exp. Bot. 2009, 67, 222–227. [Google Scholar] [CrossRef]
  146. Kausar, F.; Shahbaz, M.; Ashraf, M. Protective role of foliar-applied nitric oxide in Triticum aestivum under saline stress. Turk. J. Bot. 2013, 37, 1155–1165. [Google Scholar] [CrossRef]
  147. Ali, Q.; Daud, M.; Haider, M.Z.; Ali, S.; Rizwan, M.; Aslam, N.; Noman, A.; Iqbal, N.; Shahzad, F.; Deeba, F.; et al. Seed priming by sodium nitroprusside improves salt tolerance in wheat (Triticum aestivum L.) by enhancing physiological and biochemical parameters. Plant Physiol. Biochem. 2017, 119, 50–58. [Google Scholar] [CrossRef] [PubMed]
  148. Sehar, Z.; Masood, A.; Khan, N.A. Nitric oxide reverses glucose-mediated photosynthetic repression in wheat (Triticum aestivum L.) under salt stress. Environ. Exp. Bot. 2019, 161, 277–289. [Google Scholar] [CrossRef]
  149. Ruan, H.-H.; Shen, W.-B.; Xu, L.-L. Nitric oxide modulates the activities of plasma membrane H+-ATPase and PPase in wheat seedling roots and promotes the salt tolerance against salt stress. Acta Bot. Sin. 2004, 46, 415–422. [Google Scholar]
  150. Thomas, D.D.; Liu, X.; Kantrow, S.P.; Lancaster, J.R. The biological lifetime of nitric oxide: Implications for the perivascular dynamics of NO and O2. Proc. Natl. Acad. Sci. USA 2001, 98, 355–360. [Google Scholar] [CrossRef]
  151. García-Martínez, J. Editor Gibberellin metabolism and control of fruit growth. Int. Symp. Plant Bioregul. Fruit Prod. 1997, 463. [Google Scholar] [CrossRef] [Green Version]
  152. Gao, S.; Chu, C. Gibberellin metabolism and signaling: Targets for improving agronomic performance of crops. Plant Cell Physiol. 2020, 61, 1902–1911. [Google Scholar] [CrossRef]
  153. Sponsel, V.M.; Hedden, P. Gibberellin Biosynthesis and Inactivation. In Plant Hormones; Springer: Berlin, Germany, 2010; pp. 63–94. [Google Scholar]
  154. Serrani, J.C.; Sanjuán, R.; Ruiz-Rivero, O.; Fos, M.; García-Martínez, J.L. Gibberellin regulation of fruit set and growth in tomato. Plant Physiol. 2007, 145, 246–257. [Google Scholar] [CrossRef] [Green Version]
  155. Tudzynski, B.; Hölter, K. Gibberellin Biosynthetic Pathway inGibberella fujikuroi: Evidence for a Gene Cluster. Fungal Genet. Biol. 1998, 25, 157–170. [Google Scholar] [CrossRef]
  156. Hedden, P. The current status of research on gibberellin biosynthesis. Plant Cell Physiol. 2020, 61, 1832–1849. [Google Scholar] [CrossRef]
  157. Liu, L.; Xia, W.; Li, H.; Zeng, H.; Wei, B.; Han, S.; Yin, C. Salinity inhibits rice seed germination by reducing α-amylase activity via decreased bioactive gibberellin content. Front. Plant Sci. 2018, 9, 275. [Google Scholar] [CrossRef] [PubMed]
  158. Lv, S.; Yu, D.; Sun, Q.; Jiang, J. Activation of gibberellin 20-oxidase 2 undermines auxin-dependent root and root hair growth in NaCl-stressed Arabidopsis seedlings. Plant Growth Regul. 2018, 84, 225–236. [Google Scholar] [CrossRef]
  159. Achard, P.; Cheng, H.; De Grauwe, L.; Decat, J.; Schoutteten, H.; Moritz, T.; Straeten, D.V.D.; Peng, J.; Harberd, N.P. Integration of plant responses to environmentally activated phytohormonal signals. Science 2006, 311, 91–94. [Google Scholar] [CrossRef] [PubMed]
  160. Shan, C.; Mei, Z.; Duan, J.; Chen, H.; Feng, H.; Cai, W. OsGA2ox5, a gibberellin metabolism enzyme, is involved in plant growth, the root gravity response and salt stress. PLoS ONE 2014, 9, e87110. [Google Scholar]
  161. Zhu, N.; Cheng, S.; Liu, X.; Du, H.; Dai, M.; Zhou, D.; Yang, W.; Zhao, Y. The R2R3-type MYB gene OsMYB91 has a function in coordinating plant growth and salt stress tolerance in rice. Plant Sci. 2015, 236, 146–156. [Google Scholar] [CrossRef] [PubMed]
  162. Magome, H.; Yamaguchi, S.; Hanada, A.; Kamiya, Y.; Oda, K. The DDF1 transcriptional activator upregulates expression of a gibberellin-deactivating gene, GA2ox7, under high-salinity stress in Arabidopsis. Plant J. 2008, 56, 613–626. [Google Scholar] [CrossRef]
  163. Zhou, J.; Li, Z.; Xiao, G.; Zhai, M.; Pan, X.; Huang, R.; Zhang, H. CYP71D8L is a key regulator involved in growth and stress responses by mediating gibberellin homeostasis in rice. J. Exp. Bot. 2020, 71, 1160–1170. [Google Scholar] [PubMed]
  164. Liu, X.; Wang, X.; Yin, L.; Deng, X.; Wang, S. Exogenous application of gibberellic acid participates in up-regulation of lipid biosynthesis under salt stress in rice. Exp. Plant Physiol. 2018, 30, 335–345. [Google Scholar] [CrossRef]
  165. Wang, Y.-H.; Zhang, G.; Chen, Y.; Gao, J.; Sun, Y.-R.; Sun, M.-F.; Chen, J.-P. Exogenous application of gibberellic acid and ascorbic acid improved tolerance of okra seedlings to NaCl stress. Acta Physiol. Plant 2019, 41, 93. [Google Scholar] [CrossRef]
  166. Ahanger, M.A.; Ashraf, M.; Bajguz, A.; Ahmad, P. Brassinosteroids regulate growth in plants under stressful environments and crosstalk with other potential phytohormones. J. Plant Growth Regul. 2018, 37, 1007–1024. [Google Scholar] [CrossRef]
  167. Fariduddin, Q.; Yusuf, M.; Ahmad, I.; Ahmad, A. Brassinosteroids and their role in response of plants to abiotic stresses. Biol. Plant 2014, 58, 9–17. [Google Scholar] [CrossRef]
  168. Sirhindi, G. Brassinosteroids: Biosynthesis and Role in Growth, Development, and Thermotolerance Responses. In Molecular Stress Physiology of Plants; Springer: Berlin, Germany, 2013; pp. 309–329. [Google Scholar]
  169. Saini, S.; Sharma, I.; Pati, P.K. Versatile roles of brassinosteroid in plants in the context of its homoeostasis, signaling and crosstalks. Front. Plant Sci. 2015, 6, 950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Grove, M.D.; Spencer, G.F.; Rohwedder, W.K.; Mandava, N.; Worley, J.F.; Warthen, J.D.; Steffens, G.L.; Flippen-Anderson, J.L.; Cook, J.C. Brassinolide, a plant growth-promoting steroid isolated from Brassica napus pollen. Nature 1979, 281, 216–217. [Google Scholar] [CrossRef]
  171. Wei, Z.; Li, J. Brassinosteroids regulate root growth, development, and symbiosis. Mol. Plant 2016, 9, 86–100. [Google Scholar] [CrossRef] [Green Version]
  172. Ahammed, G.J.; Xia, X.-J.; Li, X.; Shi, K.; Yu, J.-Q.; Zhou, Y.-H. Role of brassinosteroid in plant adaptation to abiotic stresses and its interplay with other hormones. Curr. Protein Pept. Sci. 2015, 16, 462–473. [Google Scholar] [CrossRef] [PubMed]
  173. Siddiqui, H.; Ahmed, K.B.M.; Hayat, S. Comparative effect of 28-homobrassinolide and 24-epibrassinolide on the performance of different components influencing the photosynthetic machinery in Brassica juncea L. Plant Physiol. Biochem. 2018, 129, 198–212. [Google Scholar] [CrossRef]
  174. Vardhini, B.V.; Anjum, N.A. Brassinosteroids make plant life easier under abiotic stresses mainly by modulating major components of antioxidant defense system. Front. Environ. Sci. 2015, 2, 67. [Google Scholar] [CrossRef] [Green Version]
  175. Qayyum, B.; Shahbaz, M.; Akram, N.A. Interactive effect of foliar application of 24-epibrassinolide and root zone salinity on morpho-physiological attributes of wheat (Triticum aestivum L.). Int. J. Agric. Biol. 2007, 9, 584–589. [Google Scholar]
  176. Dong, Y.J.; Wang, W.W.; Hu, G.Q.; Chen, W.F.; Zhuge, Y.P.; Wang, Z.L.; He, M.R. Role of exogenous 24-epibrassinolide in enhancing the salt tolerance of wheat seedlings. J. Soil Sci. Plant Nutr. 2017, 17, 554–569. [Google Scholar] [CrossRef] [Green Version]
  177. Yusuf, M.; Fariduddin, Q.; Khan, T.; Hayat, S. Epibrassinolide reverses the stress generated by combination of excess aluminum and salt in two wheat cultivars through altered proline metabolism and antioxidants. S. Afr. J. Bot. 2017, 112, 391–398. [Google Scholar] [CrossRef]
  178. Rady, M.M. Effect of 24-epibrassinolide on growth, yield, antioxidant system and cadmium content of bean (Phaseolus vulgaris L.) plants under salinity and cadmium stress. Sci. Hortic. 2011, 129, 232–237. [Google Scholar] [CrossRef]
  179. Shahid, M.A.; Balal, R.M.; Pervez, M.A.; Garcia-Sanchez, F.; Gimeno, V.; Abbas, T.; Mattson, N.C.; Riaz, A. Treatment with 24-epibrassinolide mitigates NaCl-induced toxicity by enhancing carbohydrate metabolism, osmolyte accumulation, and antioxidant activity in Pisum sativum. Turk. J. Bot. 2014, 38, 511–525. [Google Scholar] [CrossRef]
  180. Koh, S.; Lee, S.-C.; Kim, M.-K.; Koh, J.H.; Lee, S.; An, G.; Choe, S.; Kim, S.-R. T-DNA tagged knockout mutation of rice OsGSK1, an orthologue of Arabidopsis BIN2, with enhanced tolerance to various abiotic stresses. Plant Mol. Biol. 2007, 65, 453–466. [Google Scholar] [CrossRef]
  181. Krishna, P. Brassinosteroid-mediated stress responses. J. Plant Growth Regul. 2003, 22, 289–297. [Google Scholar] [CrossRef] [PubMed]
  182. Manavalan, L.P.; Chen, X.; Clarke, J.; Salmeron, J.; Nguyen, H.T. RNAi-mediated disruption of squalene synthase improves drought tolerance and yield in rice. J. Exp. Bot. 2012, 63, 163–175. [Google Scholar] [CrossRef] [Green Version]
  183. Xia, X.-J.; Wang, Y.-J.; Zhou, Y.-H.; Tao, Y.; Mao, W.-H.; Shi, K.; Asami, T.; Chen, Z.; Yu, J.-Q. Reactive oxygen species are involved in brassinosteroid-induced stress tolerance in cucumber. Plant Physiol. 2009, 150, 801–814. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Wang, Z.-Y.; Bai, M.-Y.; Oh, E.; Zhu, J.-Y. Brassinosteroid signaling network and regulation of photomorphogenesis. Annu. Rev. Genet. 2012, 46, 701–724. [Google Scholar] [CrossRef] [PubMed]
  185. Sharma, I.; Ching, E.; Saini, S.; Bhardwaj, R.; Pati, P.K. Exogenous application of brassinosteroid offers tolerance to salinity by altering stress responses in rice variety Pusa Basmati-1. Plant Physiol. Biochem. 2013, 69, 17–26. [Google Scholar] [CrossRef]
  186. Cui, F.; Liu, L.; Zhao, Q.; Zhang, Z.; Li, Q.; Lin, B.; Wu, Y.; Tang, S.; Xie, Q. Arabidopsis ubiquitin conjugase UBC32 is an ERAD component that functions in brassinosteroid-mediated salt stress tolerance. Plant Cell 2012, 24, 233–244. [Google Scholar] [CrossRef] [Green Version]
  187. Divi, U.K.; Rahman, T.; Krishna, P. Brassinosteroid-mediated stress tolerance in Arabidopsis shows interactions with abscisic acid, ethylene and salicylic acid pathways. BMC Plant Biol. 2010, 10, 151. [Google Scholar] [CrossRef] [Green Version]
  188. Yuan, G.-F.; Jia, C.-G.; Li, Z.; Sun, B.; Zhang, L.-P.; Liu, N.; Wang, Q.-M. Effect of brassinosteroids on drought resistance and abscisic acid concentration in tomato under water stress. Sci. Hortic. 2010, 126, 103–108. [Google Scholar] [CrossRef]
  189. Bajguz, A. Brassinosteroid enhanced the level of abscisic acid in Chlorella vulgaris subjected to short-term heat stress. J. Plant Physiol. 2009, 166, 882–886. [Google Scholar] [CrossRef]
  190. Zhang, S.; Cai, Z.; Wang, X. The primary signaling outputs of brassinosteroids are regulated by abscisic acid signaling. Proc. Natl. Acad. Sci. USA 2009, 106, 4543–4548. [Google Scholar] [CrossRef] [Green Version]
  191. Gururani, M.A.; Mohanta, T.K.; Bae, H. Current understanding of the interplay between phytohormones and photosynthesis under environmental stress. Int. J. Mol. Sci. 2015, 16, 19055–19085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Hayat, S.; Ali, B.; Ahmad, A. Salicylic Acid: Biosynthesis, Metabolism and Physiological Role in Plants. In Salicylic Acid: A Plant Hormone; Springer: Berlin, Germany, 2007; pp. 1–14. [Google Scholar]
  193. Senaratna, T.; Touchell, D.; Bunn, E.; Dixon, K. Acetyl salicylic acid (Aspirin) and salicylic acid induce multiple stress tolerance in bean and tomato plants. Plant Growth Regul. 2000, 30, 157–161. [Google Scholar] [CrossRef]
  194. Tabur, S.; Avci, Z.D.; Özmen, S. Exogenous salicylic acid application against mitodepressive and clastogenic effects induced by salt stress in barley apical meristems. Biologia 2020, 76, 1–10. [Google Scholar] [CrossRef]
  195. Simaei, M.; Khavarinejad, R.; Saadatmand, S.; Bernard, F.; Fahimi, H. Interactive effects of salicylic acid and nitric oxide on soybean plants under NaCl salinity. Russ. J. Plant Physiol. 2011, 58, 783. [Google Scholar] [CrossRef]
  196. Csiszár, J.; Brunner, S.; Horváth, E.; Bela, K.; Ködmön, P.; Riyazuddin, R.; Gallé, Á.; Hurton, Á.; Papdi, C.; Szabados, L.; et al. Exogenously applied salicylic acid maintains redox homeostasis in salt-stressed Arabidopsis gr1 mutants expressing cytosolic roGFP1. Plant Growth Regul. 2018, 86, 181–194. [Google Scholar] [CrossRef]
  197. Szalai, G.; Kellős, T.; Galiba, G.; Kocsy, G. Glutathione as an antioxidant and regulatory molecule in plants under abiotic stress conditions. J. Plant Growth Regul. 2009, 28, 66–80. [Google Scholar] [CrossRef]
  198. Nazar, R.; Umar, S.; Khan, N.A. Exogenous salicylic acid improves photosynthesis and growth through increase in ascorbate-glutathione metabolism and S assimilation in mustard under salt stress. Plant Signal. Behav. 2015, 10, e1003751. [Google Scholar] [CrossRef] [Green Version]
  199. Li, G.; Peng, X.; Wei, L.; Kang, G. Salicylic acid increases the contents of glutathione and ascorbate and temporally regulates the related gene expression in salt-stressed wheat seedlings. Gene 2013, 529, 321–325. [Google Scholar] [CrossRef]
  200. Dolatabadian, A.; Modarres Sanavy, S.A.M.; Sharifi, M. Effect of salicylic acid and salt on wheat seed germination. Acta Agric. Scand. Sect. B Soil Plant Sci. 2009, 59, 456–464. [Google Scholar] [CrossRef]
  201. Kang, G.; Li, G.; Zheng, B.; Han, Q.; Wang, C.; Zhu, Y.; Guo, T. Proteomic analysis on salicylic acid-induced salt tolerance in common wheat seedlings (Triticum aestivum L.). Biochim. Biophys. Acta (BBA) Proteins Proteom. 2012, 1824, 1324–1333. [Google Scholar] [CrossRef] [PubMed]
  202. Ma, X.; Zheng, J.; Zhang, X.; Hu, Q.; Qian, R. Salicylic acid alleviates the adverse effects of salt stress on Dianthus superbus (Caryophyllaceae) by activating photosynthesis, protecting morphological structure, and enhancing the antioxidant system. Front. Plant Sci. 2017, 8, 600. [Google Scholar] [CrossRef]
  203. Yadav, T.; Kumar, A.; Yadav, R.; Yadav, G.; Kumar, R.; Kushwaha, M. Salicylic acid and thiourea mitigate the salinity and drought stress on physiological traits governing yield in pearl millet-wheat. Saudi J. Biol. Sci. 2020, 27, 2010. [Google Scholar] [CrossRef]
  204. Borsani, O.; Valpuesta, V.; Botella, M.A. Evidence for a role of salicylic acid in the oxidative damage generated by NaCl and osmotic stress in Arabidopsis seedlings. Plant Physiol. 2001, 126, 1024–1030. [Google Scholar] [CrossRef] [Green Version]
  205. Hao, L.; Zhao, Y.; Jin, D.; Zhang, L.; Bi, X.; Chen, H.; Xu, Q.; Ma, C.; Li, G. Salicylic acid-altering Arabidopsis mutants response to salt stress. Plant Soil 2012, 354, 81–95. [Google Scholar] [CrossRef]
  206. Quilis, J.; Peñas, G.; Messeguer, J.; Brugidou, C.; Segundo, B.S. The Arabidopsis AtNPR1 inversely modulates defense responses against fungal, bacterial, or viral pathogens while conferring hypersensitivity to abiotic stresses in transgenic rice. Mol. Plant Microbe Interact. 2008, 21, 1215–1231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Ding, Y.; Sun, T.; Ao, K.; Peng, Y.; Zhang, Y.; Li, X.; Zhang, Y. Opposite roles of salicylic acid receptors NPR1 and NPR3/NPR4 in transcriptional regulation of plant immunity. Cell 2018, 173, 1454–1467.e15. [Google Scholar] [CrossRef]
  208. Zhang, Y.; Li, X. Salicylic acid: Biosynthesis, perception, and contributions to plant immunity. Curr. Opin. Plant Biol. 2019, 50, 29–36. [Google Scholar] [CrossRef]
  209. Jayakannan, M.; Bose, J.; Babourina, O.; Shabala, S.; Massart, A.; Poschenrieder, C.; Rengel, Z. The NPR1-dependent salicylic acid signalling pathway is pivotal for enhanced salt and oxidative stress tolerance in Arabidopsis. J. Exp. Bot. 2015, 66, 1865–1875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Lotfi, R.; Ghassemi-Golezani, K.; Pessarakli, M. Salicylic acid regulates photosynthetic electron transfer and stomatal conductance of mung bean (Vigna radiata L.) under salinity stress. Biocatal. Agric. Biotechnol. 2020, 26, 101635. [Google Scholar] [CrossRef]
  211. Lee, S.; Kim, S.G.; Park, C.M. Salicylic acid promotes seed germination under high salinity by modulating antioxidant activity in Arabidopsis. N. Phytol. 2010, 188, 626–637. [Google Scholar] [CrossRef] [PubMed]
  212. Yuan, H.-M.; Liu, W.-C.; Lu, Y.-T. CATALASE2 coordinates SA-mediated repression of both auxin accumulation and JA biosynthesis in plant defenses. Cell Host Microbe 2017, 21, 143–155. [Google Scholar] [CrossRef]
  213. Prodhan, M.Y.; Munemasa, S.; Nahar, M.N.-E.-N.; Nakamura, Y.; Murata, Y. Guard cell salicylic acid signaling is integrated into abscisic acid signaling via the Ca2+/CPK-dependent pathway. Plant Physiol. 2018, 178, 441–450. [Google Scholar] [CrossRef] [Green Version]
  214. Montillet, J.-L.; Hirt, H. New checkpoints in stomatal defense. Trends Plant Sci. 2013, 18, 295–297. [Google Scholar] [CrossRef]
  215. Acharya, B.R.; Assmann, S.M. Hormone interactions in stomatal function. Plant Mol. Biol. 2009, 69, 451–462. [Google Scholar] [CrossRef]
  216. Zhu, Y.; Wang, B.; Tang, K.; Hsu, C.-C.; Xie, S.; Du, H.; Yang, Y.; Tao, W.A.; Zhu, J.-K. An Arabidopsis Nucleoporin NUP85 modulates plant responses to ABA and salt stress. PLoS Genet. 2017, 13, e1007124. [Google Scholar] [CrossRef] [Green Version]
  217. Wani, S.; Kumar, V. Plant stress tolerance: Engineering ABA: A Potent phytohormone. Transcriptomics 2015, 3, 1000113. [Google Scholar] [CrossRef] [Green Version]
  218. Zhang, D.-P. Abscisic Acid: Metabolism, Transport and Signaling; Springer: Berlin, Germany, 2014. [Google Scholar]
  219. Roychoudhury, A.; Paul, S.; Basu, S. Cross-talk between abscisic acid-dependent and abscisic acid-independent pathways during abiotic stress. Plant Cell Rep. 2013, 32, 985–1006. [Google Scholar] [CrossRef]
  220. Vishwakarma, K.; Upadhyay, N.; Kumar, N.; Yadav, G.; Singh, J.; Mishra, R.K.; Kumar, V.; Verma, R.; Upadhyay, R.G.; Pandey, M.; et al. Abscisic acid signaling and abiotic stress tolerance in plants: A review on current knowledge and future prospects. Front. Plant Sci. 2017, 8, 161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. Raghavendra, A.S.; Gonugunta, V.K.; Christmann, A.; Grill, E. ABA perception and signalling. Trends Plant Sci. 2010, 15, 395–401. [Google Scholar] [CrossRef] [PubMed]
  222. Balusamy, S.R.; Rahimi, S.; Yang, D.-C. Characterization of squalene-induced PgCYP736B involved in salt tolerance by modulating key genes of abscisic acid biosynthesis. Int. J. Biol. Macromol. 2019, 121, 796–805. [Google Scholar] [CrossRef] [PubMed]
  223. Hedrich, R.; Shabala, S. Stomata in a saline world. Curr. Opin. Plant Biol. 2018, 46, 87–95. [Google Scholar] [CrossRef] [PubMed]
  224. Niu, M.; Xie, J.; Chen, C.; Cao, H.; Sun, J.; Kong, Q.; Shabala, S.; Shabala, L.; Huang, Y.; Bie, Z. An early ABA-induced stomatal closure, Na+ sequestration in leaf vein and K+ retention in mesophyll confer salt tissue tolerance in Cucurbita species. J. Exp. Bot. 2018, 69, 4945–4960. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Jia, W.; Wang, Y.; Zhang, S.; Zhang, J. Salt-stress-induced ABA accumulation is more sensitively triggered in roots than in shoots. J. Exp. Bot. 2002, 53, 2201–2206. [Google Scholar] [CrossRef] [Green Version]
  226. Lu, C.; Chen, M.-X.; Liu, R.; Zhang, L.; Hou, X.; Liu, S.; Ding, X.; Jiang, Y.; Xu, J.; Zhang, J.; et al. Abscisic acid regulates auxin distribution to mediate maize lateral root development under salt stress. Front. Plant Sci. 2019, 10, 716. [Google Scholar] [CrossRef]
  227. Xing, L.; Zhao, Y.; Gao, J.; Xiang, C.; Zhu, J.-K. The ABA receptor PYL9 together with PYL8 plays an important role in regulating lateral root growth. Sci. Rep. 2016, 6, 27177. [Google Scholar] [CrossRef]
  228. Bano, A.; Aziz, N. Salt and drought stress in wheat and the role of abscisic acid. Pak. J. Bot. 2004, 35, 871–884. [Google Scholar]
  229. El-Enany, A. Abscisic acid-responsive proteins induce salinity tolerance in wheat seedlings. Acta Physiol. Plant. 2000, 22, 53. [Google Scholar] [CrossRef]
  230. Gurmani, A.; Bano, A.; Salim, M. Effect of abscisic acid and benzyladenine on growth and ion accumulation of wheat under salinity stress. Pak. J. Bot. 2007, 39, 141. [Google Scholar]
  231. Sripinyowanich, S.; Klomsakul, P.; Boonburapong, B.; Bangyeekhun, T.; Asami, T.; Gu, H.; Buaboocha, T.; Chadchawan, S. Exogenous ABA induces salt tolerance in indica rice (Oryza sativa L.): The role of OsP5CS1 and OsP5CR gene expression during salt stress. Environ. Exp. Bot. 2013, 86, 94–105. [Google Scholar] [CrossRef]
  232. Li, X.; Li, S.; Wang, J.; Lin, J. Exogenous abscisic acid alleviates harmful effect of salt and alkali stresses on wheat seedlings. Int. J. Environ. Res. Public Health 2020, 17, 3770. [Google Scholar] [CrossRef] [PubMed]
  233. Amzallag, G.; Lerner, H.; Poljakoff-Mayber, A. Exogenous ABA as a modulator of the response of sorghum to high salinity. J. Exp. Bot. 1990, 41, 1529–1534. [Google Scholar] [CrossRef]
  234. Etehadnia, M.; Waterer, D.R.; Tanino, K.K. The method of ABA application affects salt stress responses in resistant and sensitive potato lines. J. Plant Growth Regul. 2008, 27, 331. [Google Scholar] [CrossRef]
  235. Brini, F.; Hanin, M.; Lumbreras, V.; Amara, I.; Khoudi, H.; Hassairi, A.; Pagès, M.; Masmoudi, K. Overexpression of wheat dehydrin DHN-5 enhances tolerance to salt and osmotic stress in Arabidopsis thaliana. Plant Cell Rep. 2007, 26, 2017–2026. [Google Scholar] [CrossRef] [PubMed]
  236. Deng, X.; Hu, W.; Wei, S.; Zhou, S.; Zhang, F.; Han, J.; Chen, L.; Li, Y.; Feng, J.; Fang, B.; et al. TaCIPK29, a CBL-interacting protein kinase gene from wheat, confers salt stress tolerance in transgenic tobacco. PLoS ONE 2013, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Huang, X.; Zhang, Y.; Jiao, B.; Chen, G.; Huang, S.; Guo, F.; Shen, Y.; Huang, Z.; Zhao, B. Overexpression of the wheat salt tolerance-related gene TaSC enhances salt tolerance in Arabidopsis. J. Exp. Bot. 2012, 63, 5463–5473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Nemoto, Y.; Sasakuma, T. Differential stress responses of early salt-stress responding genes in common wheat. Phytochemistry 2002, 61, 129–133. [Google Scholar] [CrossRef]
  239. Yang, T.; Yao, S.; Hao, L.; Zhao, Y.; Lu, W.; Xiao, K. Wheat bHLH-type transcription factor gene TabHLH1 is crucial in mediating osmotic stresses tolerance through modulating largely the ABA-associated pathway. Plant Cell Rep. 2016, 35, 2309–2323. [Google Scholar] [CrossRef]
  240. Ambastha, V.; Friedmann, Y.; Leshem, Y. Laterals take it better–emerging and young lateral roots survive lethal salinity longer than the primary root in Arabidopsis. Sci. Rep. 2020, 10, 1–11. [Google Scholar] [CrossRef] [PubMed]
  241. Roberts, S.K. Regulation of K+ channels in maize roots by water stress and abscisic acid. Plant Physiol. 1998, 116, 145–153. [Google Scholar] [CrossRef] [Green Version]
  242. Zaid, A.; Mohammad, F. Methyl jasmonate and nitrogen interact to alleviate cadmium stress in Mentha arvensis by regulating physio-biochemical damages and ROS detoxification. J. Plant Growth Regul. 2018, 37, 1331–1348. [Google Scholar] [CrossRef]
  243. Ali, M.; Baek, K.-H. Jasmonic acid signaling pathway in response to abiotic stresses in plants. Int. J. Mol. Sci. 2020, 21, 621. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Arif, Y.; Singh, P.; Siddiqui, H.; Bajguz, A.; Hayat, S. Salinity induced physiological and biochemical changes in plants: An omic approach towards salt stress tolerance. Plant Physiol. Biochem. 2020, 156, 64–77. [Google Scholar] [CrossRef] [PubMed]
  245. Kang, D.J.; Seo, Y.J.; Lee, J.D.; Ishii, R.; Kim, K.U.; Shin, D.H.; Park, S.K.; Jang, S.W.; Lee, I.J. Jasmonic acid differentially affects growth, ion uptake and abscisic acid concentration in salt-tolerant and salt-sensitive rice cultivars. J. Agron. Crop Sci. 2005, 191, 273–282. [Google Scholar] [CrossRef]
  246. Pedranzani, H.; Racagni, G.; Alemano, S.; Miersch, O.; Ramírez, I.; Peña-Cortés, H.; Taleisnik, E.; Domenech, E.-M.; Abdala, G. Salt tolerant tomato plants show increased levels of jasmonic acid. Plant Growth Regul. 2003, 41, 149–158. [Google Scholar] [CrossRef]
  247. Zhang, H.; Zhang, Q.; Zhai, H.; Li, Y.; Wang, X.; Liu, Q.; He, S. Transcript profile analysis reveals important roles of jasmonic acid signalling pathway in the response of sweet potato to salt stress. Sci. Rep. 2017, 7, 1–12. [Google Scholar] [CrossRef] [Green Version]
  248. Zhao, Y.; Dong, W.; Zhang, N.; Ai, X.; Wang, M.; Huang, Z.; Xiao, L.; Xia, G. A wheat allene oxide cyclase gene enhances salinity tolerance via jasmonate signaling. Plant Physiol. 2014, 164, 1068–1076. [Google Scholar] [CrossRef] [Green Version]
  249. Valenzuela, C.E.; Acevedo-Acevedo, O.; Miranda, G.S.; Vergara-Barros, P.; Holuigue, L.; Figueroa, C.R.; Figueroa, P.M. Salt stress response triggers activation of the jasmonate signaling pathway leading to inhibition of cell elongation in Arabidopsis primary root. J. Exp. Bot. 2016, 67, 4209–4220. [Google Scholar] [CrossRef] [Green Version]
  250. Sheteawi, S.A. Improving growth and yield of salt-stressed soybean by exogenous application of jasmonic acid and ascobin. Int. J. Agric. Biol. 2007, 9, 473–478. [Google Scholar]
  251. Ali, A.Y.A.; Ibrahim, M.E.H.; Zhou, G.; Nimir, N.E.A.; Jiao, X.; Zhu, G.; Elsiddig, A.M.I.; Suliman, M.S.E.; Elradi, S.B.M.; Yue, W. Exogenous jasmonic acid and humic acid increased salinity tolerance of sorghum. Agron. J. 2020, 112, 871–884. [Google Scholar] [CrossRef]
  252. Yastreb, T.O.; Kolupaev, Y.E.; Shkliarevskyi, M.A.; Dyachenko, A.I.; Dmitriev, A.P. Jasmonate signaling components’ participation in stomata closing induced by salt sress in Arabidopsis thaliana. Tsitol. Genet. 2020, 54, 46–52. [Google Scholar]
  253. Li, C.; Lv, J.; Zhao, X.; Ai, X.; Zhu, X.; Wang, M.; Zhao, S.; Xia, G. TaCHP: A wheat zinc finger protein gene down-regulated by abscisic acid and salinity stress plays a positive role in stress tolerance. Plant Physiol. 2010, 154, 211–221. [Google Scholar] [CrossRef] [Green Version]
  254. Wang, M.; Xia, G. The landscape of molecular mechanisms for salt tolerance in wheat. Crop J. 2018, 6, 42–47. [Google Scholar] [CrossRef]
  255. Jiang, Y.; Deyholos, M.K. Comprehensive transcriptional profiling of NaCl-stressed Arabidopsis roots reveals novel classes of responsive genes. BMC Plant Biol. 2006, 6, 25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Ma, S.; Gong, Q.; Bohnert, H.J. Dissecting salt stress pathways. J. Exp. Bot. 2006, 57, 1097–1107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Kilian, J.; Whitehead, D.; Horak, J.; Wanke, D.; Weinl, S.; Batistic, O.; D’Angelo, C.; Bauer, E.B.; Kudla, J.; Harter, K. The AtGenExpress global stress expression data set: Protocols, evaluation and model data analysis of UV-B light, drought and cold stress responses. Plant J. 2007, 50, 347–363. [Google Scholar] [CrossRef]
  258. Geng, Y.; Wu, R.; Wee, C.W.; Xie, F.; Wei, X.; Chan, P.M.Y.; Tham, C.; Duan, L.; Dinneny, J.R. A spatio-temporal understanding of growth regulation during the salt stress response in Arabidopsis. Plant Cell 2013, 25, 2132–2154. [Google Scholar] [CrossRef] [Green Version]
  259. Ebel, C.; BenFeki, A.; Hanin, M.; Solano, R.; Chini, A. Characterization of wheat (Triticum aestivum) TIFY family and role of Triticum Durum TdTIFY11a in salt stress tolerance. PLoS ONE 2018, 13, e0200566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  260. Khan, M.I.R.; Khan, N.A. Ethylene reverses photosynthetic inhibition by nickel and zinc in mustard through changes in PS II activity, photosynthetic nitrogen use efficiency, and antioxidant metabolism. Protoplasma 2014, 251, 1007–1019. [Google Scholar] [CrossRef]
  261. Chang, C. Q&A: How do plants respond to ethylene and what is its importance? BMC Biol. 2016, 14, 7. [Google Scholar]
  262. Lei, G.; Shen, M.; Li, Z.G.; Zhang, B.; Duan, K.X.; Wang, N.; Cao, Y.-R.; Zhang, W.K.; Ma, B.; Ling, H.-Q.; et al. EIN2 regulates salt stress response and interacts with a MA3 domain-containing protein ECIP1 in Arabidopsis. Plant Cell Environ. 2011, 34, 1678–1692. [Google Scholar] [CrossRef] [PubMed]
  263. Thao, N.P.; Khan, M.I.R.; Thu, N.B.A.; Hoang, X.L.T.; Asgher, M.; Khan, N.A.; Tran, L.-S.P. Role of ethylene and its cross talk with other signaling molecules in plant responses to heavy metal stress. Plant Physiol. 2015, 169, 73–84. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Müller, M.; Munné-Bosch, S. Ethylene response factors: A key regulatory hub in hormone and stress signaling. Plant Physiol. 2015, 169, 32–41. [Google Scholar] [CrossRef] [Green Version]
  265. Sun, X.; Zhao, T.; Gan, S.; Ren, X.; Fang, L.; Karungo, S.K.; Wang, Y.; Chen, L.; Li, S.; Xin, H. Ethylene positively regulates cold tolerance in grapevine by modulating the expression of ethylene response factor 057. Sci. Rep. 2016, 6, 1–14. [Google Scholar] [CrossRef] [PubMed]
  266. Arraes, F.B.M.; Beneventi, M.A.; de Sa, M.E.L.; Paixao, J.F.R.; Albuquerque, E.V.S.; Marin, S.R.R.; Purgatto, E.; Nepomuceno, A.L.; Grossi-de-Sa, M.F. Implications of ethylene biosynthesis and signaling in soybean drought stress tolerance. BMC Plant Biol. 2015, 15, 213. [Google Scholar] [CrossRef] [PubMed]
  267. Habben, J.E.; Bao, X.; Bate, N.J.; DeBruin, J.L.; Dolan, D.; Hasegawa, D.; Helentjaris, T.G.; Lafitte, R.H.; Lovan, N.; Mo, H.; et al. Transgenic alteration of ethylene biosynthesis increases grain yield in maize under field drought-stress conditions. Plant Biotechnol. J. 2014, 12, 685–693. [Google Scholar] [CrossRef] [PubMed]
  268. Pierik, R.; Tholen, D.; Poorter, H.; Visser, E.J.; Voesenek, L.A. The Janus face of ethylene: Growth inhibition and stimulation. Trends Plant Sci. 2006, 11, 176–183. [Google Scholar] [CrossRef]
  269. Zapata, P.J.; Serrano, M.; García-Legaz, M.F.; Pretel, M.; Botella, M. Short term effect of salt shock on ethylene and polyamines depends on plant salt sensitivity. Front. Plant Sci. 2017, 8, 855. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  270. Morgan, P.W.; Drew, M.C. Ethylene and plant responses to stress. Physiol. Plant 1997, 100, 620–630. [Google Scholar] [CrossRef]
  271. Li, C.-H.; Wang, G.; Zhao, J.-L.; Zhang, L.-Q.; Ai, L.-F.; Han, Y.-F.; Sun, D.-Y.; Zhang, S.-W.; Sun, Y. The receptor-like kinase SIT1 mediates salt sensitivity by activating MAPK3/6 and regulating ethylene homeostasis in rice. Plant Cell 2014, 26, 2538–2553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  272. Xu, J.; Li, Y.; Wang, Y.; Liu, H.; Lei, L.; Yang, H.; Liu, G.; Ren, D. Activation of MAPK kinase 9 induces ethylene and camalexin biosynthesis and enhances sensitivity to salt stress in Arabidopsis. J. Biol. Chem. 2008, 283, 26996–27006. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Yang, L.; Zu, Y.-G.; Tang, Z.-H. Ethylene improves Arabidopsis salt tolerance mainly via retaining K+ in shoots and roots rather than decreasing tissue Na+ content. Environ. Exp. Bot. 2013, 86, 60–69. [Google Scholar] [CrossRef]
  274. Freitas, V.S.; de Souza Miranda, R.; Costa, J.H.; de Oliveira, D.F.; de Oliveira Paula, S.; de Castro Miguel, E.; Freire, R.S.; Prisco, J.T.; Gomes-Filho, E. Ethylene triggers salt tolerance in maize genotypes by modulating polyamine catabolism enzymes associated with H2O2 production. Environ. Exp. Bot. 2018, 145, 75–86. [Google Scholar] [CrossRef]
  275. Shakeri, S.; Saeid, N.F.; Safipour, A.A. Effect of salicylic acid and ethephon on seed germination and seedling growth of wheat under salt stress. J. Crop. Ecophysiol. (Agric. Sci.) 2016, 10, 779–790. [Google Scholar]
  276. Cao, W.-H.; Liu, J.; He, X.-J.; Mu, R.-L.; Zhou, H.-L.; Chen, S.-Y.; Zhang, J.-S. Modulation of ethylene responses affects plant salt-stress responses. Plant Physiol. 2007, 143, 707–719. [Google Scholar] [CrossRef] [Green Version]
  277. Riyazuddin, R.; Verma, R.; Singh, K.; Nisha, N.; Keisham, M.; Bhati, K.K.; Kim, S.T.; Gupta, R. Ethylene: A master regulator of salinity stress tolerance in plants. Biomolecules 2020, 10, 959. [Google Scholar] [CrossRef]
  278. Lockhart, J. Salt of the earth: Ethylene promotes salt tolerance by enhancing Na/K homeostasis. Am. Soc. Plant Biol. 2013. [Google Scholar] [CrossRef] [Green Version]
  279. Peng, J.; Li, Z.; Wen, X.; Li, W.; Shi, H.; Yang, L.; Zhu, H.; Guo, H. Salt-induced stabilization of EIN3/EIL1 confers salinity tolerance by deterring ROS accumulation in Arabidopsis. PLoS Genet. 2014, 10, e1004664. [Google Scholar] [CrossRef] [Green Version]
  280. Lang, T.; Deng, C.; Yao, J.; Zhang, H.; Wang, Y.; Deng, S. A salt-signaling network involving ethylene, extracellular ATP, hydrogen peroxide, and calcium mediates K+/Na+ homeostasis in Arabidopsis. Int. J. Mol. Sci. 2020, 21, 8683. [Google Scholar] [CrossRef] [PubMed]
  281. Egamberdieva, D. Alleviation of salt stress by plant growth regulators and IAA producing bacteria in wheat. Acta Physiol. Plant 2009, 31, 861–864. [Google Scholar] [CrossRef]
  282. Xu, Z.-S.; Xia, L.-Q.; Chen, M.; Cheng, X.-G.; Zhang, R.-Y.; Li, L.-C.; Zhao, Y.-X.; Lu, Y.; Ni, Z.-Y.; Liu, L.; et al. Isolation and molecular characterization of the Triticum aestivum L. ethylene-responsive factor 1 (TaERF1) that increases multiple stress tolerance. Plant Mol. Biol. 2007, 65, 719–732. [Google Scholar] [CrossRef] [PubMed]
  283. Rong, W.; Qi, L.; Wang, A.; Ye, X.; Du, L.; Liang, H.; Xin, Z.; Zhang, Z. The ERF transcription factor TaERF 3 promotes tolerance to salt and drought stresses in wheat. Plant Biotechnol. J. 2014, 12, 468–479. [Google Scholar] [CrossRef] [PubMed]
  284. Djemal, R.; Khoudi, H. TdSHN1, a WIN1/SHN1-type transcription factor, imparts multiple abiotic stress tolerance in transgenic tobacco. Environ. Exp. Bot. 2016, 131, 89–100. [Google Scholar] [CrossRef]
  285. Jang, C.S.; Lee, H.J.; Chang, S.J.; Seo, Y.W. Expression and promoter analysis of the TaLTP1 gene induced by drought and salt stress in wheat (Triticum aestivum L.). Plant Sci. 2004, 167, 995–1001. [Google Scholar] [CrossRef]
  286. Gangadhar, B.H.; Sajeesh, K.; Venkatesh, J.; Baskar, V.; Abhinandan, K.; Yu, J.W.; Prasad, R.; Mishra, R.J. Enhanced tolerance of transgenic potato plants over-expressing non-specific lipid transfer protein-1 (StnsLTP1) against multiple abiotic stresses. Front. Plant Sci. 2016, 7, 1228. [Google Scholar] [CrossRef] [Green Version]
  287. Xu, Y.; Zheng, X.; Song, Y.; Zhu, L.; Yu, Z.; Gan, L.; Zhou, S.; Liu, H.; Fujian, W.; Zhu, C. NtLTP4, a lipid transfer protein that enhances salt and drought stresses tolerance in Nicotiana tabacum. Sci. Rep. 2018, 8, 1–14. [Google Scholar] [CrossRef] [Green Version]
  288. Hu, W.; Yuan, Q.; Wang, Y.; Cai, R.; Deng, X.; Wang, J.; Zhou, S.; Chen, M.; Chen, L.; Huang, C.; et al. Overexpression of a wheat aquaporin gene, TaAQP8, enhances salt stress tolerance in transgenic tobacco. Plant Cell Physiol. 2012, 53, 2127–2141. [Google Scholar] [CrossRef] [Green Version]
  289. Ma, Q.; Shi, C.; Su, C.; Liu, Y. Complementary analyses of the transcriptome and iTRAQ proteome revealed mechanism of ethylene dependent salt response in bread wheat (Triticum aestivum L.). Food Chem. 2020. [Google Scholar] [CrossRef]
  290. Aldrich, S. Epibrassinolide 2021. Available online: https://www.sigmaaldrich.com/catalog/product/sigma/e1641?lang=en&region=AU&cm_sp=Insite-_-caSrpResults_srpRecs_srpModel_epibrassinolide-_-srpRecs3-1 (accessed on 20 May 2021).
  291. Smallbon, T. Plant Density, Spacing and Health Underpin Green-Pea Returns. 2018. Available online: https://www.utas.edu.au/__data/assets/pdf_file/0004/1130368/Pea-Project-Fact-Sheet-Season-2-2018.pdf (accessed on 20 May 2021).
  292. FAO. Producer Prices 2020. Available online: http://www.fao.org/faostat/en/#data/PP (accessed on 20 May 2021).
Figure 1. Effect of cytokinin on salt stress tolerance in plants (based on ❶ [115,130], ❷ [131], ❸ [127,132], ❹ [133], ❺ [134], and ❻ [128,129]).
Figure 1. Effect of cytokinin on salt stress tolerance in plants (based on ❶ [115,130], ❷ [131], ❸ [127,132], ❹ [133], ❺ [134], and ❻ [128,129]).
Biomolecules 11 00788 g001
Figure 2. Effect of salicylic acid on salt stress tolerance in plants (based on ❶ [212], ❷ [213], ❸ [214,215], ❹ [52], ❺ [193,200], ❻ [52,207], and ❼ [207]).
Figure 2. Effect of salicylic acid on salt stress tolerance in plants (based on ❶ [212], ❷ [213], ❸ [214,215], ❹ [52], ❺ [193,200], ❻ [52,207], and ❼ [207]).
Biomolecules 11 00788 g002
Figure 3. Effect of abscisic acid on salt stress tolerance in plants (based on ❶ [240], ❷ [223], ❸ [241], ❹ [236], ❺ [231], and ❻ [239]).
Figure 3. Effect of abscisic acid on salt stress tolerance in plants (based on ❶ [240], ❷ [223], ❸ [241], ❹ [236], ❺ [231], and ❻ [239]).
Biomolecules 11 00788 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Quamruzzaman, M.; Manik, S.M.N.; Shabala, S.; Zhou, M. Improving Performance of Salt-Grown Crops by Exogenous Application of Plant Growth Regulators. Biomolecules 2021, 11, 788. https://doi.org/10.3390/biom11060788

AMA Style

Quamruzzaman M, Manik SMN, Shabala S, Zhou M. Improving Performance of Salt-Grown Crops by Exogenous Application of Plant Growth Regulators. Biomolecules. 2021; 11(6):788. https://doi.org/10.3390/biom11060788

Chicago/Turabian Style

Quamruzzaman, Md., S. M. Nuruzzaman Manik, Sergey Shabala, and Meixue Zhou. 2021. "Improving Performance of Salt-Grown Crops by Exogenous Application of Plant Growth Regulators" Biomolecules 11, no. 6: 788. https://doi.org/10.3390/biom11060788

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop